You are on page 1of 82

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Review

Cite This: Chem. Rev. XXXX, XXX, XXX-XXX pubs.acs.org/CR

Photochemistry and Photophysics in Silica-Based Materials: Ultrafast


and Single Molecule Spectroscopy Observation
Noemí Alarcos,† Boiko Cohen,† Marcin Ziółek,‡ and Abderrazzak Douhal*,†

Departamento de Química Física, Facultad de Ciencias Ambientales y Bioquímica, and INAMOL, Universidad de Castilla-La
Mancha, Avenida Carlos III, S.N., 45071 Toledo, Spain

Quantum Electronics Laboratory, Faculty of Physics, Adam Mickiewicz University, Umultowska 85, 61-614 Poznań, Poland

ABSTRACT: Silica-based materials (SBMs) are widely used in catalysis, photonics, and
drug delivery. Their pores and cavities act as hosts of diverse guests ranging from
classical dyes to drugs and quantum dots, allowing changes in the photochemical
behavior of the confined guests. The heterogeneity of the guest populations as well as
the confinement provided by these hosts affect the behavior of the formed hybrid
materials. As a consequence, the observed reaction dynamics becomes significantly
different and complex. Studying their photobehavior requires advanced laser-based
spectroscopy and microscopy techniques as well as computational methods. Thanks to
the development of ultrafast (spectroscopy and imaging) tools, we are witnessing an
increasing interest of the scientific community to explore the intimate photobehavior of
these composites. Here, we review the recent theoretical and ultrafast experimental
studies of their photodynamics and discuss the results in comparison to those in
homogeneous media. The discussion of the confined dynamics includes solvation and
intra- and intermolecular proton-, electron-, and energy transfer events of the guest
within the SBMs. Several examples of applications in photocatalysis, (photo)sensors, photonics, photovoltaics, and drug delivery
demonstrate the vast potential of the SBMs in modern science and technology.

CONTENTS 3.3.2. Energy Transfer in Dye-Doped Mesopo-


rous Materials AF
1. Introduction B 3.3.3. Donor/Acceptor Concentration Effects
2. Theoretical Studies D on Energy Transfer AI
2.1. Solvation in Silica-Based Materials E 3.3.4. Energy Transfer in Silica-Coated Quan-
2.1.1. Water and Silica Materials E tum Dots and Metal Nanoparticles AJ
2.1.2. Other Solvents and Silica Materials F 3.4. Homo-Energy Transfer and Aggregates
2.1.3. Impact on Solvation Dynamics G Formation AL
2.2. Photoinduced Processes in Trapped Mono- 4. Single Molecule AP
mers and Aggregates I 4.1. Single Molecule Reactivity in Silica-Based
2.2.1. Impact of SBMs on Absorption and Materials AQ
Emission Spectra of Chromophores I 4.2. Single Molecule Diffusion in Silica-Based
2.2.2. Impact on Excited-State Dynamics J Materials AR
2.2.3. Impact of Guest Aggregation K 4.3. Other Studies AR
3. Ensemble Average Time-Resolved Studies of 4.4. Electronic Nanoconfinement AS
Photoinduced Processes in Silica-Based Materials L 5. Recent Applications of Silica-Based Materials AT
3.1. Proton-Transfer Reactions L 5.1. Photocatalysis AT
3.1.1. Intermolecular Proton-Transfer Pro- 5.2. Photonics AY
cesses M 5.2.1. Sensors BA
3.1.2. Intramolecular Proton-Transfer Reac- 5.3. Photovoltaics BE
tions S 5.4. Drug Delivery BG
3.2. Electron-Transfer Reactions U 6. General Conclusion and Outlook BJ
3.2.1. Intramolecular Charge-Transfer Reac- Author Information BK
tions V Corresponding Author BK
3.2.2. Intermolecular Electron-Transfer Reac- ORCID BK
tions AA
3.3. Energy-Transfer Events AD
3.3.1. Energy Transfer in Dye-Doped Zeolites AE
Received: July 13, 2017

© XXXX American Chemical Society A DOI: 10.1021/acs.chemrev.7b00422


Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 1. Frameworks of the Silica-Based Materials (SBMs) Used in the Studies Commented in This Reviewa

a
(A) and (B) show the microporous and mesoporous hosts, respectively.

Notes BK guest dynamics and spectroscopy different from those in


Biographies BK solutions.2,7,44−49
Acknowledgments BK In recent years, many studies of ensemble average ultrafast
Dedication BK (femtosecond regime) processes within the confined space
Abbreviations BK provided by the nanochannels and nanocavities of the zeolite
General Abbreviations BK and mesoporous materials have been published. However, the
Symbols and Rate Constants BL most recent systematic review of photodynamical studies using
Material Structures BL SBM composites was published in 2003.4 Thus, there is a
Molecules BM strong need for a detailed review of the recent contributions on
References BN the photoinduced fast and ultrafast dynamical properties of
such SBMs interacting with chromophores and drugs. Here, by
fast and ultrafast dynamics we mean events happening in pico-
1. INTRODUCTION and femtosecond (ps-fs) regimes. The present review mainly
Silica-based materials (SBMs) are one of the most studied focuses on contributions since 2005. However, in a few places,
framework systems in advanced chemical and biological earlier relevant contributions are also presented. We also review
applications, being used in catalysis, photonics, and drug recent works using time-resolved fluorescence microscopy
delivery. They act as hosts for diverse organic molecules (dyes applied to SBM composites. From the point of view of the
and drugs for example), allowing changes in the photochemical framework materials, our interest is focused on microporous
behavior of the confined guests. The formed composites show a silica and aluminosilica sieves (e.g., zeolites), mesoporous silica
variety of photophysical and photochemical events that could sieves (e.g., MCM41, SBA15), silica nanoparticles, and silica
be used to develop smart devices, such as drug nanocarriers, shells. Scheme 1 shows the framework of these SBMs confining
nanosensors, nanoOLEDs, nanolasers, energy storage nano- media, which have been used in the published works
space, and nanophotocatalysts. commented in this review. As for the encapsulated
Several review articles on the structural properties of SBMs chromophores, the main part is devoted to organic molecules,
and their composites with different chromophores have been but quantum dots (QDs), metal nanoparticles, and perovskites
published.1−17 However, these works have not been oriented are also presented. The chemical structure of these guests is
toward the photodynamics in the formed composites. They shown in Schemes 2−8.
rather focus on the photochemistry at long time scale Theoretical studies of the photophysical and photochemical
(nanosecond-millisecond regime). Therefore, a great knowl- processes in silica-based composites are, without doubt, much
edge on slow photochemical events and stability of the formed more difficult in comparison to the calculations of isolated
photoproducts has been reached and used for some application molecules or even those in a homogeneous environment.
purposes.3,5,13−15,18−28 More than two decades ago, ultrafast Caution must be taken when interpreting the results, due to the
laser-based techniques allowed one to witness the chemical use of more approximated methods or base sets. Similarly, time-
bond and electronic events in real-time of reactions.29,30 These resolved experiments in such hybrid systems become a
techniques are being used to characterize the molecular events challenge against conventional studies in solutions because of
in condensed phase, providing details on solvation, electron, their more complex and less repeatable samples preparation,
proton, and energy transfer events. 16,27,31−43 Confining difficulties in successful measurements (e.g., studies in solid
molecules within chemical and biological cavities leads to state or in suspensions where the light is severely scattered),
B DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 2. Molecular Structures of the Compounds Interacting with SBMs and Discussed in the Theoretical Section 2

and often a more complicated data analysis. We believe that the streak-cameras, and fluorescence microscopy), the possibility of
time of ultrafast studies of simple molecular systems dissolved studying complex systems without having a detailed knowledge
in solutions is passing. Such systematic studies in the past 30 in ultrafast laser physics is growing. An analogous situation
years resulted in profound knowledge about the photophysical happens in the field of theoretical calculations, where the rapid
processes that occur in solutions on the time scales of development of computational tools has empowered the field in
femtoseconds (fs), picoseconds (ps), and nanoseconds (ns). the two last decades.
The rapid increase in novel technologies requires the scientists In addition to the Introduction and Conclusion, this review
to focus more on systems that are application-oriented, contains four sections. The review is organized as follows: the
including the composites formed by chromophores in SBMs.
subsequent section (section 2) is devoted to the theoretical
The basic studies of novel chromophores in solutions are still
necessary (e.g., to get to know the effects of surrounding studies. Its first section 2.1 examines the ultrafast dynamics of
polarity, viscosity, pH, and specific interactions) but rather as water and other solvents within the pores and cages provided
references to investigate more complex environments. With the by the SBMs, the solvation effects on the solutes, and its impact
recent development of turn-key ultrafast laser systems of high on the observed dynamics and spectroscopy. The second
stability, and commercially available time-resolved spectrom- section of the theoretical part (section 2.2) presents the studies
eters (e.g., transient absorption in UV, visible, NIR ranges, up- of the electronic coupling between the guest and the framework
conversion, time-correlated single-photon counting (TCSPC), of the host and its relevance to the photobehavior of the
C DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

formed composites. The effects of aggregation and energy in areas such as photocatalysis, photonics, sensors, photo-
transfer between closely laying guests are also considered. voltaics, and drug delivery. A few results of ns−ms
Section 3 (ensemble average time-resolved studies of luminescence (that were not presented in parts 3−4) related
photoinduced processes in SBMs) is the main and the longest to application-oriented studies are also commented in section 5.
section of the review, presenting and discussing the recent The review ends with the conclusions that can be drawn on the
reports on the dynamics and spectroscopy of different basis of the photophysical processes within SBMs, and we
photoinduced processes taking place within silica-based hosts address future challenges and the possibility of how to face
on the time scale from fs to ns. Section 3.1 deals with the them.
excited-state proton-transfer reactions that can be either intra The goal that this review, among others, is to provide a guide
or intermolecular in nature. We review the effects that the of general changes in the photodynamics (with respect to
silica-based hosts have on these processes. The framework of homogeneous solvents) that one can expect when considering a
the host can affect directly or indirectly the excited-state given chromophore interacting with silica-based hosts and
proton-transfer reactions by modifying the electronic distribu- nanostructures. For example, excited guests, which have a
tion within the guest and, thus, altering its dynamics and significant change in their dipole moments, should exhibit
subsequent events. In specific cases, the framework can act as a slower solvation (nanosolvation) with respect to that in
proton acceptor/donor, inducing an intermolecular proton- solutions, and the extent of this slowing down depends on
transfer reaction within the composites. The effects of the pore their position and adsoption on the silica pores and framework
size, metal doping, and functionalization of the host on the governing the specific and nonspecific interactions in the
proton-transfer event of the guest are also discussed. Section composites. When the guest nonradiative deactivation depends
3.2 is devoted to electron-transfer reactions (one of the most on its internal rotation or twisting, we anticipate that its
important reactions in nature), and it includes both the passive fluorescence lifetime in the composite will increase as a result of
and active role of the host. For the first case, the host restriction in motion. On the other hand, other processes can
frameworks can affect the charge separated (CS) states of the make the photodynamics faster in SBMs than those in
encapsulated guest, together with the rate constants of intra- solutions. Excited state lifetime might become shorter due to
and intermolecular charge transfer reactions, and the nano- self-quenching (homo energy transfer) upon high dye loading
solvation dynamics. As for the active role, framework in the pores, and the size of host pores might favor or disfavor
modification using metal doping functionalization or pore size aggregates formation. On the other hand, metal doping of silica
can further affect the charge separation (within the guest) and materials can trigger additional deactivation channels, due to
can even facilitate electron injection to the doping metal. We charge transfer (e.g., electron injection). Small modifications in
also examine the effects of specific and nonspecific interactions acidity or basicity inside silica pores might influence the proton-
between the guest and the host on the electron-transfer events. transfer reaction rates and promote the deactivation route at
SBMs are one of the most suitable hosts to build multi- the expense of another one. Often, it may even lead to the
chromophoric scaffolds for energy transfer, and the related occurrence of a new molecular form (e.g., radical anions,
reports dedicated to understanding these events and developing cations, and CS state) in the ground state, which drastically
multichromophoric and hybrid systems by functionalization, changes the overall photophysics and photochemistry of the
doping, and pore size modification are commented on in confined system. These are the most common and simplest
section 3.3. The effects of energy donor and acceptor examples, with much more details, special cases, and competing
concentrations in the nanohosts on the processes in silica- factors outlined in the following parts.
coated QDs and metal nanoparticles are considered as well. Our review of the fs−ps time-resolved studies of the
Finally, section 3.4 analyzes the results from the point of view photoinduced processes taking place in the hybrid SBM
of molecular aggregates formation. The limited space of the complexes should give a concise perspective on the clues and
nanochannels and nanocavities of the zeolites and mesoporous knowledge of the challenges found in the photocharacterization
materials creates physical conditions for strong intermolecular of these composites. One of its goals is to trigger further
interactions between the encapsulated guest molecules, thus research to advance in the related subfields and emerge with
giving rise to H- and J-aggregates of the caged monomers. We others in science and technology. We hope that the review will
discuss how this type of guest−guest interactions shapes the be of great interest to a wide scientific community (in
photophysical behavior of the composites, and how doping and chemistry, nanomaterials, spectroscopy, physics, nanomedicine,
functionalization of the host can affect the distribution of the and pharmacology) that will benefit from having at hand the
aggregates and interacting monomers. state-of-the-art in the ultrafast photodynamics within silica-
Section 4 of the review shows the most recent studies based composites, encouraging more and more scientific groups
oriented toward using the photophysical properties of the to join this fascinating and promising field to boost the
guests to characterize the hybrid complexes at single molecule/ knowledge opening doors for new smart applications of SBM
particle level. Knowing the interaction of fluorescent dyes with composites.
the silica-based host framework is paramount to deciphering
the distribution of the guests and the type of specific and 2. THEORETICAL STUDIES
nonspecific interactions in the composites, by passing the There is a strong discrepancy between the number of
ensemble average limitations. Special attention is paid to the theoretical and experimental works that examine ultrafast
electronic nanoconfinement, reflecting the effect of the host photoinduced processes in silica-based materials (SBMs). To
framework on the photophysical properties of the guest from predict the behavior of the system after light absorption, the
the point of view of orbitals coupling. Section 5 summarizes the interaction of the excited state of the organic molecules with
most relevant applications of SBM composites that are related silica-based structures must be optimized, which is difficult not
to the photophysical and photochemical processes presented in only because of the large systems that are involved but also due
sections 2−4. Thus, we focused on the applications using light to the lack of proper methodology, which is still being
D DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

developed (even in the case of small molecules). Another shell model) divides solvent (water) molecules into two
problem is the verification of the calculated results: the groups: (1) bound (surface) and (2) free (bulk) water. Surface,
calculations are usually performed on small molecules, while or shell, molecules are assumed to have significantly modified
the experiments, involving dyes on SBMs, are usually properties due to their direct interaction with silica and their
performed with large molecules possessing the desired immobilization on its surface via H-bonding. In contrast, core
properties and generating strong enough signals to analyze molecules distant from the interface are considered to have
(e.g., high extinction coefficient in the visible range, strong bulklike properties. A given property of a confined liquid
dipole moment changes upon excitation, etc.). should scale as 1/R and 1/R2 in cylindrical and spherical
Compared with the case just described, there are relatively confinements, respectively, due to the populations of the
many theoretical reports that address ground-state interactions molecules in the bound and free states (R is the radius of the
and adsorption of gases and small liquid molecules on SBMs. confining space). A more complex model assumes that the
These studies, particularly relevant to catalysis, have been influence of the interface extends further and more smoothly
recently reviewed.15 Another area of theoretical studies of into the liquid than the step function used in a two-state model
adsorption and diffusion of small molecules in mesoporous describes. A dynamic exchange model attributes the long-time
silica is devoted to sensing application (see section 5). For dynamics to the exchange of bound and free water (a bound
example, a physico-mathematical model for the related kinetics water molecule can only take part in solvation dynamics by
was recently proposed to predict the kinetic behavior of the exchanging with a free water molecule). This model leads to the
explosive trinitrotoluene (TNT) in mesoporous films.50 On the prediction of a temperature dependence of the solvation
contrary, studying the adsorption of large biomolecules is dynamics that is due to an activation energy caused by a free
important in, for example, drug delivery (see section 5). Such energy barrier. This barrier is generated by the H-bonding
calculations have been recently commented on.11 The above character of water molecules with the silica surface. Finally, a
reviews also outlined the main theoretical methods that are model with chromophore diffusion assumes that changes occur
used to get the structures of SBMs and their interactions with in the position of the solute molecule when its dipole moment
the organic guest molecules.11,15 Therefore, these subjects will changes (excitation). We will discuss this model in the next
not be considered in this review. Below, we mainly focus on the section.
theoretical works that are dedicated to ultrafast and fast More recently, the reorientation dynamics of water in
dynamics in SBMs and the properties of the nanoconfined nanoscale hydrophilic or hydrophobic silica pores were
molecules in their excited states or, at least, those that predict investigated using molecular dynamics simulations.51 The
their stationary absorption and emission spectra. The first reorientation of water in hydrophilic confinement occurs in
section is devoted to solvation dynamics that influences the more than one nanosecond (2 orders of magnitude longer than
relaxation of the excited states of the dyes. In the second in bulk water) and exhibits a power-law decay. The hydro-
section, the direct interaction of the dyes with SBMs will be phobic pores (simulated as silica with no charges) affect more
considered with regards to photoinduced processes. modestly the dynamics, making water molecules reorientation
only 2.5 times slower than that in the bulk. Moreover, the
2.1. Solvation in Silica-Based Materials
power-law decay of the correlation function in the hydrophilic
Solvation dynamics is determined by the ability of solvent pores was explained in terms of heterogeneity of the surface
molecules to reorganize around a chromophore, thereby and demonstrates the incompleteness of the two-state model.
minimizing the energy of the resulting solvent−solute entity. The H-bond jump model was used to study the heterogeneity
The typical and most frequently used way to study solvation where H-bond acceptors on the pore surface (bridging oxygens
dynamics is the electronic excitation of a molecule and gating or silanols) exchange H-bonding interactions with water
its fluorescence signal at short and different times after its molecules. This mechanism, when combined with the surface
excitation. A suitable chromophore should have different dipole roughness, leads to a distribution of jump times. Obviously,
moments in the ground and excited states in a polar solvent. If water molecules cannot form H-bonds with the interface when
the electronically first excited state (S1) has significant they are in the hydrophobic pores, and the reorientation
intramolecular charge transfer (ICT) character, solvent dynamics is governed only by the local excluded volume.
molecules around the solute move and rotate to assume the Molecular dynamics simulations combined with analytic
energetically optimum position. The stabilization of the excited modeling were also used to explain why water reorientation
solute in the new electronic configuration is then accompanied dynamics does not vary monotonically with the surface
by a redshift in its fluorescence spectra (Stokes shift). This shift hydrophilicity.52 The increasing hydrophilic character is
with the gating time is analyzed to get solvation time dynamics. accompanied by two competing effects within the interfacial
2.1.1. Water and Silica Materials. An important result layer. The first effect is a rearrangement of water molecules. In
from the studies of solvation dynamics is predicting the this scenario, when the hydrophilicity of a surface increases, an
interactions of solvent molecules with SBMs, particularly in increasing fraction of water molecules point their OH groups
materials in which nanoconfinement occurs. Such studies will toward the surface. Further increases in hydrophilicity increases
be briefly discussed below, while the experimental findings are the interaction energy, slowing down the reorientation times.
presented in section 3.2.1. Not surprisingly, water is the most It is well-established that while fully hydroxylated surfaces
frequently theoretically studied solvent because of its simple (surface density of 4−5 Si−OH groups per nm2) have a
molecular structure, well-known strong interactions with SBMs, hydrophilic character, highly dehydroxylated ones (approx-
and relevance in biology and medicine, and to our life. imately 1−2 Si−OH per nm2) manifest a hydrophobic
Advances in the field using theoretical tools were reviewed in character.11 The extended jump model in cases of hydrophilic
2011.33 Different models describing the dynamical behavior of confinement was further studied by spatially resolving the H-
solvent molecules in confined structures were summarized bond jump dynamics among individual sites on the silica pore
there. The simplest model (the so-called two-state or core− surface.53 The nonexponential dynamics was explained by a
E DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 1. Picture of the molecular dynamics of water in MCM41, after (A) 3.5 ps (potential energy equilibration reached) and (B) 6.5 ps. The water
shell (blue) is almost identical at both times. Reprinted with permission from ref 58. Copyright 2016 Royal Society of Chemistry.

broad distribution of rate constants for the exchange of H- correlation of the spectra at a distance of 2 Å, even until 1 ps.
bonding partners, and the distribution was due to a variety of While at a distance of 5 Å, the OH groups lose the memory of
local topographies at the silica surface. It was found that the key their frequency much faster (although still slower than in the
factors that determine the distribution of jump types were bulk liquid). Therefore, the theoretical report shows that it
entropic: specifically, an excluded-volume effect for the should be possible to probe the slower spectral diffusion of
approach of a new partner and the elongation of the initial confined water molecules (compared to the bulk liquid) by
H-bond. However, enthalpy effects arising from the chemical performing and analyzing 2D-IR spectra.
heterogeneity of the pore surface contributed only weakly to Another recent work used density functional theory (DFT)
the distribution. The same research group investigated the based-molecular dynamics simulations to study water interact-
hydrophobic confinement in all-silica Linde type A (LTA) ing with SBMs. Amorphous silica surfaces and quartz surfaces
zeolites.54 Water molecules reorientation was found to be were compared when they were in contact with water.57 It was
retarded by only a factor of 2−3 when compared with the bulk concluded that the water layer at the amorphous silica interface
values. The slowdown of water dynamics was more pronounced is less dipolar than the one at the crystalline silica interface. The
at higher water loadings and was predominantly caused by an layer of water molecules at the amorphous silica surface was
excluded-volume effect (large fraction of water molecules lying found to be immobile, and an adsorbed water molecule would
at the interface within the zeolite matrix). The authors probably not be easily replaced by adsorbate organic molecules.
suggested that the presence of the interface and its chemical Other composite models at the DFT level with explicit solvent
nature have a much greater impact on water dynamics than water molecules were tested for the classical Mobil Catalytic
does the confinement. Materials of Number 41 (MCM41).58 Here, ab initio molecular
Classical molecular dynamics simulations were used to study dynamics using an MCM41 channel having a silanol density of
the interaction of water with the external surface of silicalite-1.55 5.8 Si−OH nm−2 showed that the first water layer is strongly
Despite the hydrophobic behavior of the silicalite structure and adsorbed on the silica wall (blue in Figure 1). The mobility of
the presence of hydrophilic OH groups on the surface, water the first shell of water is qualitatively very low, and water
enters the zeolite pore system and forms a layer. The adsorbed molecules stay close to the channel surface and do not diffuse
water molecules on this internal layer are H-bonded to each during sampling times of up to 8 ps (Figure 1).
other and to a layer of molecules that are adsorbed on the 2.1.2. Other Solvents and Silica Materials. The
surface and bound to the silanol groups. The surface layer has interaction of nonaqueous solvents with SBMs was also
an ordered structure with a thickness of approximately 7 Å. theoretically studied. Molecular dynamics simulations of a
The performed calculations were not only aimed at silica/acetonitrile interface were performed.59 Acetonitrile
explaining experimental findings but also predicting new results molecules formed an arrangement at the silica surface that is
and phenomena. Molecular dynamics calculations were recently reminiscent of a lipid bilayer. Such a configuration is
used to simulate the results of a two-dimensional infrared considerably different from an antiparallel and dipole-paired
spectroscopy (2D-IR) experiment, in which a novel ultrafast structure in the bulk liquid. It was also found that the silica
technique probed the correlation and coherence times of the interface can affect the bulk liquid structure over a distance of
molecular vibrations.56 The results predict that while the tens of angstroms. A similar bilayer structure was considered at
frequency correlation of OH vibrations is nearly complete in the interface between silica and propionitrile.60 However, the
approximately 2 ps for bulk water, in a silica pore, this CN vector of the cyanide group of propionitrile is more parallel
correlation is significantly longer (6 ps). Moreover, the 2D to the surface than that of acetonitrile. As a result, the
description of the distance of OH vibrations from the pore propionitrile bilayer is more compact and ordered than the
surface drastically changed: there was little loss in the frequency corresponding acetonitrile one (Figure 2A).
F DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 2. (A) Comparison of the CN (vector of cyanine group) density (ρ) in acetonitrile and propionitrile (z is the center of the CN bond, ρB is
the bulk density. Reprinted from ref 60. Copyright 2012 American Chemical Society. (B) Number of hydrogen bonds per methanol molecule as a
function of distance from the silica surface and resultant polarity of the environment. Reprinted from ref 61. Copyright 2013 American Chemical
Society.

Figure 3. (A) Variation of free energy profiles, ΔA(x), and probability distributions, P(x), of a solute of different dipole moments: 5 D (black), 10 D
(blue), and 15 D (green) in the hydrophilic pore. The figure shows that the solute of a larger dipole moment (e.g., in the excited state) is located
further from the pore walls. Reprinted from ref 64. Copyright 2012 American Chemical Society. (B) A model system with a spherical solute in a
Lennard-Jones fluid confined between a slit of parallel walls at the distance H, below the solvation time obtained for solvophilic and solvophobic
walls (solute diameter is 3σ). Reprinted from ref 65. Copyright 2013 American Chemical Society.

Another theoretical report showed that the Coulombic methanol molecules that were dissolved in acetonitrile and
interaction of the π-electrons of benzene ring with the partial confined in nanoscale hydrophilic silica pores.63 Methanol
charges carried by the silica atoms is of great importance in the molecules with specific interactions were found to exist in two
stability of the complex.62 The confined benzene exhibits distinct conformations that differ in their H-bonding states,
significant layering, and the rings of molecules close to the which resulted in two maxima in the position-dependent
interface tend to lie flat on the silica surface. The rotational densities of the molecules with respect to the distance from the
dynamics of this solvent in the center of the pore was found silica surface. Both entropy as well as H-bonding were found to
slightly slower when compared to that of the bulk, with more play important roles in determining the location and
pronounced slowing in smaller pores. For example, the orientation of the methanol molecules.
reorientation time at 293 K was predicted to be ∼6 and ∼7 2.1.3. Impact on Solvation Dynamics. The configu-
ps in nanopores with diameters of 3.6 and 2.0 nm, respectively, rations of solvent molecules and their reorientation times that
while at the same temperature, this time was predicted ∼5 ps are commented above have a direct effect on the solvation
for bulk benzene. dynamics of chromophores that interact with silica surfaces in
Molecular dynamics simulations were also used to character- the presence of solvent. Experimental observations from time-
ize methanol.61 Strong H-bonds between the first layer of dependent fluorescence measurements (that were supported by
methanol and the silanol groups of silica create a methyl- theoretical calculations) were divided into three groups: (i)
terminated surface that results in a second layer having increase in solvation times observed in bulk solution, (ii)
significantly less density and H-bonds than the bulk solution appearance of new, longer times in the solvation dynamics, and
(Figure 2B). In agreement with the experimental findings, the (iii) lack of any differences among the chromophores in bulk
interfacial environment in this layer resembles to a nonpolar solutions and those within silica structures.33 The used
solvent. Moreover, solvent reorientation times, in these first chemical models (SBMs interacting with a solvent or a solute)
two layers, are significantly slower than those observed in the before 2010 that were proposed to explain the experimental
bulk solvent. Recently, replica exchange molecular dynamics observations for nanoconfined silica structures have also been
simulations were used to investigate the location of acetone and summarized.33 It was argued that the chemical models with
G DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 4. (A) Coumarin 153 (C153) in methanol confined in different silica pores (from left to right: HOC, HIC, and RHOC). (B) Simulation of
changes in solvation-response function, S(t), with the time. The inset shows the unnormalized energy gaps, ΔE(t), which represents the environment
contribution (i.e., MeOH and pore walls), to the energy gap between the S1 and S0 Born−Oppenheimer potential energy surfaces of C153 at time t.
Reprinted from ref 67. Copyright 2011 American Chemical Society.

different solvent layers were insufficient to accurately reproduce restriction of solvent dynamics and faster in-plane solvent
the experimental results, and that the effects of chromophore diffusion.
diffusion upon excitation must be considered. This diffusion More recently, studies of coumarin 153 (C153, Scheme 2) in
occurs when the dipole moment of the chromophore increases ethanol and confined in 2.4 nm hydrophilic amorphous silica
as a result of the electronic distribution changes from the pore have been reported.66 At the ground state, the most likely
ground to the electronically excited state, and this diffusion location of C153 when interacting with the SBMs is near the
should be responsible for the long-time solvation dynamics. In pore surface, but two possible H-bonding structures lead to
addition to that, the restriction of the solvent molecules to different orientations. Internal energy and entropy competed
move at the interface decreases its polarity relative to the host within the pore. The implication for solvation dynamics is that
interior. Therefore, the dependence of the location of a solute the two calculated orientations might lead to diffusions of
molecule (guest) on its electronic charge distribution during different rates upon excitation. These differences in rates are
the time getting the photoresponse of the composite, and the because the flat orientation of C153 has a larger effective
gradient in effective solvent polarity are likely important surface area (slower diffusion) to move into the interior than
features to take into account in understanding the behavior the second, perpendicular orientation. In a subsequent work,
of these systems. the positions of C153 in a hydroxyl-terminated silica pore
The solute (guest) diffusion model has been further used. A system was studied in both the ground and excited states.66 The
model molecule that was solvated by ethanol in a nanoscale predicted movement of C153 toward the pore interior upon
silica pore was examined as a function of its position (Figure excitation was attributed to more favorable nonspecific ethanol
3A).64 The guest distribution depends on both its dipole solvation of the large dipole moment in the excited C153,
moment and its interaction with the surface of the silica pore, which occurred at the expense of H-bonding with the pore.
which can be hydrophilic or hydrophobic in nature. The Interestingly, the free energy variations were calculated to be
electrostatic guest-pore interactions were responsible for the comparatively small (1−3 kcal/mol), which suggests that
shifts in the fluorescence spectrum of the guest molecules in relatively modest changes in the confining framework or solute
some locations near the hydrophilic wall. However, this effect interactions may generally have a significant impact on the
was absent when the guest is in the hydrophobic pore. distribution of solute positions.
Therefore, while the guest-position distributions depend on its Solvation response functions of C153 were also simulated
dipole moment in both hydrophilic and hydrophobic pores, the from averages of non-equilibrium independent trajectories
fluorescence spectra are only sensitive to the dye position in the obtained in molecular dynamic calculations.67 The investigated
former case. As a result, time-dependent fluorescence measure- solvent was methanol, and three kinds of silica pores were
ments of the hydrophobic pore system could be insufficient to examined (Figure 4A): hydrophobic cavities (HOC, in which
observe guest diffusion after excitation. However, the state- wall-solvent interactions were exclusively of the dispersive,
dependent on its position might still have significant Lennard-Jones type), hydrophilic cavities (HIC, in which
importance in applications of porous materials. unsaturated oxygen sites at the wall were transformed into OH
The differences between hydrophilic and hydrophobic groups), and rugged pores (RHOC, in which 60% of the polar
confinement were also studied using another approach.65 The groups were transformed into trimethylsilyl moieties). The
authors generalized the case of water to any solvent, using the results of the simulation within the first 30 ps are presented in
terms of solvophilic and solvophobic probe. A substantial Figure 4B. The overall responses were found to be between 2
slowing of solvation dynamics around a solute in strong and 4.5 times slower than the one observed in the bulk
solvophilic confinement is due to the suppression of fluid methanol, the one associated with RHOC being the fastest
diffusion in the presence of the closely lying attractive walls, in relaxation and the one corresponding to HIC the slowest.
addition to restricted solvent dynamics. The solvation in strong Finally, the kinetic processes occurring on the ps−ns time
solvophobic confinement is slower than in the bulk but is not scale were also considered in the calculations of diffusion of
slowed as significantly as in the solvophilic case (Figure 3B). In different molecules within SBMs. Molecular dynamics simu-
this case, the effect was explained by the competition between lations were performed and visualized on pure silica zeolite
H DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

silicalite MFI framework. The microporous zeolite contains two Similar conclusions were drawn from calculating the
channel systems that intersect tangentially, straight channels preferred positions of FL in zeolite L.73 As commented upon
and zigzag or sinusoidal channels. The diffusion dynamics were above, the overall redshift of features in the absorption
simulated by modeling trapping in the subsequent cavities spectrum in hydrated zeolite L from their positions in
separated by thermally activated jumps through the chan- cyclohexane samples was found to originate mainly from the
nels.68,69 These simulations were performed maintaining both interactions between the counterions of the zeolite and the
the framework flexibility and the realistic account of electro- carbonyl group of FL. It was found that the direct effect of
static interactions with adsorbed water. Two types of water molecules on the excitation energies of the dye is
adsorption, specific and weak unspecific, were predicted on negligible, but they do affect its orientation inside the zeolite
the channel walls and at the channel intersection. The channel. The role of water molecules in Ibu interactions with an
molecular diffusion of water in the silicalite found a barrier amorphous silica surface was also studied.74 Simulations
for crossing between the straight and the zigzag channels. Next, revealed that the silica surface exhibits a higher affinity for
the analysis of the thermal motion predicted that at room water than for Ibu. Van der Waals forces are the leading
temperature, framework oxygen atoms incurring into the zeolite contributor to its adsorption, independently of whether the
channels significantly influence the dynamics of adsorbed drug is H-bonded directly to the surface or via water molecules.
water.69 The effect of immobilization of cationic Ir(III) complexes on
Further calculations were performed for ibuprofen (Ibu, sodium aluminosilicate mesoporous glass was experimentally
Scheme 2) release from pH-gated silica nanochannels.70 The and theoretically studied.75 The calculations considered the
Ibu delivery process from cylindrical silica pores of 3 nm chemical environment of the Ir (III) complex by including both
diameter, with polyamine chains anchored at the pore outlets, the counterion (representing the active site of the matrix) and
was investigated by means of massive molecular dynamics the dielectric constant of the matrix. At low dielectric constants
simulations. High, low, and intermediate pH environments (ε < 20), the counterion has a strong influence on the energy
were investigated. It was reported that the increment of the and orbitals of the excited states. The influence of zeolite Y
acidity of the environment leads to a significant decrease of the framework on a nickel(II) complex was also examined.76 The
pore aperture, yielding an effective diameter, for the lowest pH deviations that occur in the geometry of the complex upon
case, that is 3.5 times smaller than the one associated with the encapsulation result in changes of the highest occupied
highest pH one. The joint analysis of the corresponding Gibbs molecular orbital and lowest unoccupied ones (HOMO and
free energy profiles for the Ibu delivery process, the time LUMO respectively) energy levels, and in the electronic
evolution of its position within the channel, orientation of the properties of the system. The calculated chemical hardness
molecule, and instantaneous effective diameter of the gate confirmed that the encapsulated nickel(II) complex is more
suggest a three-step mechanism for Ibu delivery. A comple- stable than the free one, and it can undergo more easily
mentary analysis of its translational mobility along the axial electron transfer (ET) reactions, thus performing photo-
direction of the channel revealed a subdiffusive dynamics in the catalysis better than the free one.
low and intermediate pH cases.70 The calculations of absorption spectra of organic molecules
2.2. Photoinduced Processes in Trapped Monomers and interacting with SBMs were performed for possible applications
Aggregates in catalysis. For example, (poly)aromatic cationic compounds
In this section, results of theoretical predictions on the have characteristic absorption bands in the visible region, and
interactions of dyes with the SBMs will be commented on. the benefits of using a theoretical approach to identify them as
First, we examine the effects of the interaction with the silica potential candidates was suggested.77 The origin of their
environment on the absorption and emission spectra. Then, we absorption spectra has been attributed to structurally different
address its impact on the excited state lifetimes and finish with species on the basis of calculated excitation energies. Moreover,
the reports on interguests interaction that leads to energy it was shown that the use of time-dependent density functional
transfer (EnT) between nearby trapped dyes within SBMs pore theory (TD-DFT) computations on gas-phase compounds was
or cavity. insufficiently accurate.78 It was suggested that molecular
2.2.1. Impact of SBMs on Absorption and Emission dynamic simulations are necessary to consider geometrical
Spectra of Chromophores. When trapped fluorenone (FL, deformations of carbonaceous compounds as these lead to
Scheme 2) in the channels of zeolite L was theoretically broad absorption bands: the results of such simulations were in
investigated, the interaction of its carbonyl group with the excellent agreement with the experimental data.
potassium cations of the zeolite framework was found to be DFT/TD-DFT calculations were also used to explain the
responsible for its stabilization in the composites.71 The result of experiments on t-azobenzene in Linde type L zeolites
calculations predicted optical anisotropy in this composite, (t-Ab and LTL in Scheme 2 and Scheme 1, respectively).79 In
maintained upon contact with water, which can drastically alter contrast to its planar structure in solution, t-Ab is twisted in
the preferred position of the dye in the silica nanostructure. LTL zeolites, as it is constrained by the framework. The
Such effects were calculated too for oxonine and pyronine photophysical properties in acidic (H-LTL) and potassium (K-
cationic dyes in zeolite L (Ox+ and Py+, respectively, Scheme LTL) forms of the zeolites were explained on the basis of
2).72 Different preferred dye molecule orientations were (n,π*) and (π,π*) transitions of the dye. In K-LTL zeolites,
obtained in the presence or absence of water, as a result of both transitions are present in the absorption spectra, with
stabilization of either the host or the guest. In the hydrated intensity maxima slightly red-shifted from their positions in
composite, a perpendicular orientation of the Ox+ or Py+ solution samples. In contrast, in H-LTL zeolite, the (n,π*)
molecules with respect to the zeolite channel axis was favored, transition was not observed as the lone electrons pair of the
while the order of preferred orientations was reversed under nitrogen atom forms a H-bond with a Brønsted acid site in the
dry conditions. zeolite. These results confirmed that t-Ab was protonated in H-
I DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 5. (A) Molecular structure of DEAC molecule showing the dihedral angle, θ. (B) Ground and excited state potential energy and oscillator
strength as a function of θ. (C) Population density distributions as a function of time and θ. Reprinted with permission from ref 85. Copyright 2014
Elsevier B.V.

LTL, and that only this species can fluoresce, as it does not absorption and emission spectra of the incorporated dyes were
connect with the dark (n,π*) state. red-shifted compared to the spectra in toluene. C339 and C340
In addition to organic chromophores as guests, absorption molecules were further studied theoretically by comparing their
spectra of silica-encapsulated quantum dots (QDs) were also photobehavior in ethanol and in Pluronic silica nanoparticles.84
investigated.80 Strong coupling between Cd2Te2 QDs and silica Two molecules from that family were investigated: one of them
shell was predicted, leading to the distortion of the silica with flexible alkyl moieties and the other one has rigid alkyl
nanocage. Therefore, a direct electronic transition from the groups. Electronic excitation results in a twisted-intramolecular
occupied Cd2Te2 states to the outer silica nanocage excited charge-transfer (TICT) state in the former molecule, which
states (core-shell electronic transitions) was predicted. As a leads to a short fluorescence decay in solutions that is different
result, the absorption peak of Cd2Te2 absorption spectrum is from that of the rigid molecule. However, both molecules
shifted from 584 nm (isolated) to 534 nm when the QDs is showed similar photophysical behaviors when embedded into
encapsulated by silica. In section 3.3, we comment on the the nanoparticles, which was shown to be governed by the
experimental observations and the involved EnT processes. drastically hampered deactivation of the flexible molecule in the
2.2.2. Impact on Excited-State Dynamics. Tetrame- silica matrix.84 Redshift in emission spectra (dye in MCM41 vs
thylrhodamine isothiocyanate (TRITC, Scheme 2) in water and toluene solution) and their temperature effects were calculated
covalently encapsulated to silica nanoparticle was studied by a in a subsequent work by the same group.85 The authors used
combination of TD-DFT and classical molecular mechanic TD-DFT with a generalized Smoluchowski equation approach
simulations, in addition to fluorescence spectroscopy.81 The to compute time-resolved spectra of 7-diethylaminocoumarin-
authors studied the changes in the molecular structure upon 3-carboxylic acid butylamine ester dye free in ethanol and
electronic excitation and found that the variations were indeed encapsulated in silica nanoparticles (DEAC, Scheme 2 and
smaller for the encapsulated dye than in water. A subsequent Figure 5A). The predicted emission quenching occurred
report explained the reason for the long excited state lifetime of because of the conformational change from an ICT state to a
TRITC in silica nanoparticles.82 It was found that the crossing TICT state (Figure 5). The rotation that leads to this change
from (π,π*) to (n,π*) state in water solutions is reduced by the was significantly hindered when the dye interacted with silica.
rigid environment of the silica shell, decreasing the fluorescence The theoretically predicted restrictions in rotational motions of
quenching of the bright (π,π*) state by the dark (n,π*) one. encapsulated molecules found confirmation in experimental
Two 7-aminocoumarin dyes that were incorporated into reports (see section 3).
MCM41 were studied using DFT and TD-DFT calculations Possible ET reactions from the excited states of carotenoids
(coumarin 339 and 340, C339 and C340, respectively, Scheme to the mesoporous frameworks of MCM41 and its derivatives
2).83 It was shown that the interaction of the carbonyl groups of were studied, being of possible applications in dye-sensitized
these dyes with the silanol groups on the silica surface is solar cells. Sections 3 and 5 comment on the involved events
responsible for stabilizing them in the composites and is and their applications in photovoltaics. It was found that the
strengthened upon photoexcitation. As a result, the computed formation of H-bonds stabilizes the canthaxanthin (CAN,
J DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 6. (A) Kohn−Sham molecular orbitals of the dimer of 7-aminocoumarin. Reprinted with permission from ref 84. Copyright 2013 Royal
Society of Chemistry. (B) Illustration of the distribution of guest (green bars) in a homogeneous (right) and inhomogeneous way (left) inside a
zeolite L crystal for high loadings. Their impact on FRET dynamics is highlighted by the different number of transfer steps (red circles) expected in a
trajectory (indicated as black lines). The blue lines show the distribution of guests along the channels. Reprinted from ref 93. Copyright 2016
American Chemical Society.

Scheme 2) molecule in CuMCM41 and affects its charge a monomer to the binding LUMO of the dimer leads to a
distribution and LUMO energy.86 As a result, CAN exhibits a strong interaction between the two molecules. Interestingly, the
lower photoinduced ET efficiency than β-carotene, for which results address two commonly observed dynamic properties of
no H-bond between the dye and the host is formed. This effect nanoconfined chromophores that have opposite effects. On one
was further studied in a subsequent work by the same group, hand, blocking the twisting of a molecule encapsulated in SBMs
predicting two types of H-bonds between the OH group of can lead to a longer excited state lifetime. On the other hand,
retinol and silanol ones on the surface of MCM41.87 One of the interaction of guests with the silica surface may induce the
these H-bonds decreases the LUMO energy of the interacting formation of their aggregates, which accelerates the deactivation
carotenoid and stabilizes its neutral species more than its radical of the excited state due to the EnT between the bright and the
cation, thus disfavoring photoinduced ET from the carotenoid dark states in the aggregates.
to MCM41. The opposite behavior was found for the second The orientation and photophysical properties of methyl-
type of H-bonds. acridine dye in zeolite L were calculated (MeAcr+, Scheme 2).94
2.2.3. Impact of Guest Aggregation. Increasing the The most stable orientation of this guest results in both its long
concentration of a guest in the silica framework usually leads to and short molecular axes being nearly perpendicular to the
aggregates formation. Formation of these entities is closely channel axis and was found mainly determined by dye-zeolite L
related to the occurrence of EnT between nearby adsorbed electrostatic interactions in addition to that with the cosolvent,
guest molecules. EnT from photoexcited guest (donors) to water. The optimized conformations were used to explain the
acceptor molecules behaves according to the model of Förster observed EnT processes (self-absorption). FRET in dye-filled
resonance EnT (FRET).88 In particular, the fluorescence monolithic crystals of zeolite L was recently modeled using
kinetics for such a process should be described by a Monte Carlo and molecular dynamic simulations.93 Under ideal
stretched-exponential function (sometimes also called Kohl- conditions (homogeneous dye distribution), very high exciton
rausch decay function).88,89 The equation has the following diffusion rates were expected in zeolite L. However, the
formula: f (t) = A exp [(−t/τs)α], where τs is a characteristic inhomogeneity of the dye distribution along with the 1D
time of the decay and α is the heterogeneity parameter (α = 1 channels led to a large change in the exciton dynamics that
for monoexponential decay, and a smaller value indicates more reduces the exciton lifetimes and modifies the dimensionality of
dispersive kinetics and large sample heterogeneity).88,90−92 For the energy transport (Figure 6B).
the 7-aminocoumarin interacting with SBMs discussed FRET was also studied in LTL zeolites and interpreted with
previously, the effect of the dye concentration in a matrix on the help of theoretical calculations.95 Cyanine (1,1′-diethyl-2,2′-
the fluorescence quenching was also theoretically explored.84 cyanine iodide, PIC) and acriflavine hydrochloride (AF) dyes
The calculations indicated that the strong quenching is acted as energy donor and acceptor partners, respectively
attributable to the formation of excimers with a CT character (Scheme 2). Energy was transferred from excited PIC to AF
upon excitation of the monomeric species. The LUMO of the that were adsorbed into the internal surfaces of zeolite LTL. A
calculated aggregate (dimer) has a strong bonding character redshift of AF emission band in the protonated LTL zeolite was
(Figure 6A). The excitation of an electron from the HOMO of explained in terms of a conversion from a cationic to a
K DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 3. Molecular Structures of the Compounds Interacting with SBMs and Discussed in the Proton-Transfer Section 3.1

protonated AF form. The experimental reports on EnT in 3. ENSEMBLE AVERAGE TIME-RESOLVED STUDIES OF
PHOTOINDUCED PROCESSES IN SILICA-BASED
SBMs, and its application are highlighted, respectively, in
MATERIALS
section 3.3, and section 5.
3.1. Proton-Transfer Reactions
In this part, we have commented on how theoretical studies
Proton-transfer (PT) reactions are intensively investigated
have become gradually important tools in explaining and since they play important roles in a wide variety of chemical
and biological processes, such as acid−base reactions,
predicting the short-time dynamics in SBMs. A large part of biochemistry, and catalysis (see Scheme 3 for molecules
theoretical research is devoted to modeling solvation dynamics involved in PT reactions discussed here).32,37−39,96−107 More-
over, the study of these reactions has resulted in the
in silica-based hosts and the resulting effects of different local development of many applications, such as fluorescent
chemosensors,108,109 laser dyes,110−114 ultraviolet (UV) photo-
environment at different positions in the silica pores. Other
stabilizers,115 photoswitches,116−119 and organic optoelectronic
studies predicted how the interaction of the chromophore with materials.120−122
These reactions occur between proton (or hydrogen atom)
the host changes its photophysical properties and how it affects donor (typically −OH, −NH2, and −SH) and acceptor (CO,
the stationary absorption/emission spectra and the photo- −N, base and some solvents) groups. The reaction can be
inter- or intramolecular and can be activated thermally or by
dynamics upon light excitation. light; the process can occur through or without a
L DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

barrier.32,97,101,105,107 Aromatic molecules that contain an acidic f em t o s e c o n d t i m e s c a le i n g a s a n d c o n d e n se d


proton in their structures can undergo intermolecular PT phases.29,32,40,100,106,149,157,165−170 Coupled proton and electron
reactions in both the ground (ground-state proton transfer, motions in the molecular system, of great importance in many
GSPT) and the electronically excited (excited-state proton chemical and biological systems, have been experimentally and
transfer, ESPT) states.37,98,101,123,124 Upon electronic excitation, theoretically examined.34,171−173 The experimental and the-
these molecules become strong acids (photoacids) that can oretical findings not only generate new knowledge but also
undergo an ESPT reaction with a neighboring base molecule, open new research fields with possible applications in sensing,
such as water, alcohols, or amines.39,102,123−127 Since Weller’s catalysis, drug delivery, photovoltaics, and photonics (section
and Eigen’s works,96,128−130 the mechanisms of PT reactions 5).
have been intensively investigated from experimental and Both inter- and intramolecular PT dynamics are sensitive to a
theoretical points of view.100,101,107,124,127,131−133 The Smolu- variety of environmental properties, such as polarity, hydro-
chowski-Collins-Kimball (SCK) model has been used to phobicity, viscosity, acidity, basicity, and H-bonding abil-
describe this type of reactions.134−137 This model shows how ity.117,120 This sensitivity to the surrounding characteristics
the diffusion of the acid and the base through the solvent plays makes PT systems potential candidates as environmental
an important role in the mechanism, as the reactants should sensors, an important aspect that has been explored by several
approach each other at a specific distance for the PT reaction to groups.32,38,117,120 Researchers took advantage of this virtue by
occur. The surrounding environment may play a key factor in using molecules that transfer protons to characterize the nano-
solute/solvent interactions and subsequent reorganization. and microenvironments in cavities and pores of SBMs. In these
Therefore, the global dynamics of these reactions are systems, the specific and nonspecific molecular interactions
determined by diffusion, interactions with the solvent, and between the trapped guest (the proton-transfer dye) and the
the adiabatic and nonadiabatic nature of the potential energy host, as well as with the caged solvent molecules, can be
surfaces (PES), as well as by the type of the prototropic modified to tune the spectroscopy and dynamics of the formed
groups.123,124,138−141 composites (dye/SBMs).
In bifunctional organic molecules that contain both the 3.1.1. Intermolecular Proton-Transfer Processes. In the
proton donor and acceptor groups in close proximity, an presence of bases (solvent or another solute), many molecules
intramolecular PT reaction in the ground (GSIPT) or that have acidic protons undergo intermolecular PT processes
electronically excited (ESIPT) state can occur.32,99,117,142 in both S0 and S1 states (GSPT and ESPT, respec-
Because of the chemical structure and the positions of the tively).98,123,124,174−178 Central to this review, most of the
partners in these bifunctional systems, an intramolecular H- SBMs have OH groups, and thus are interesting material
bond (IHB) is formed in the ground state (S0), usually models to study the PT reaction dynamics of guest molecules
producing a species called an enol (E) form. Consequently, within their pores or channels.
upon electronic excitation of E, an intramolecular electronic 3.1.1.1. Molecules Exhibiting Deprotonation with the
charge redistribution occurs in E* that leads to the formation of Framework in S0 and S1 States. Upon encapsulation of
the keto (K*) or zwitterion (Z*) tautomer. When these species chromophores in SBMs, they can establish an intermolecular
return to S0, they can be stabilized in this or other tautomeric H-bond with the surrounding framework; protonation or
forms (e.g., rotated K), or return to the E species.143−148 The
deprotonation in the ground and/or excited state are possible
mechanism and rate constant of the ESIPT reactions depend
events. 2-Naphthol (2-NpOH, Scheme 3) is a classic example
on the PESs of the E* and K* tautomers in the first
that demonstrates a PT reaction with a H-bonded acceptor
electronically first excited single state (S1).101,149 The presence
group (water, alcohols, amines, etc.). 2-NpOH emission in
or absence of barriers for both processes in S0 and S1 opens the
water solutions is dominated by its anion’s (2-naphtholate)
possibility of controlling the related spectroscopy and
emission as a result of a large change in the acidity of 2-NpOH
dynamics.44,150−155 Thus, for a PES without an energy barrier
between the wells of E* and K* species, the ESIPT reaction is in S1 (pKa = 9.5 and pKa* = 2.8).124,179−182 Its spectroscopy
ultrafast (fs regime) and irreversible.101,156 When the PES has a and related dynamics are pH-dependent.183 In acidic and
barrier in the conversion of E* to K*, the reaction may be neutral media (pH = 1−7), which have relatively high [H+], the
reversible or irreversible but will be slower (ps−ns regime) than ESPT reaction is described by a reversible mechanism (Figure
in the barrierless situation.34,157−164 An accurate description of 7). In this accepted scheme, kPT and krec are the deprotonation
the reaction dynamics at the PESs of ESIPT processes in gas and recombination rate constants, while kD and k−D are the
phase requires at least two reaction coordinates: (i) the proton forward and reverse diffusion-controlled rate constants that can
motion within the corresponding IHB and (ii) the vibrational be obtained from the Debye-Smoluchowski equation.124,183,184
or torsional (angle) mode involving the distance between the In contrast, in alkaline conditions (pH > 7) and in the presence
partners; the angle between the involved moieties in the of sodium acetate, the ESPT reaction moves forward and is
transfer could be important in the global reaction dynamics and
subsequent processes.97,99,101,107 In the presence of an energy
barrier in the PESs, the proton motion may occur via tunneling,
which adds another (quantum) dimension to the reaction. In
solution, the PESs might be complex, involving a solvent
coordinate where its polarity and H-bonding could play an
important role in shaping the spectroscopy and dynamics. The
earliest studies of proton-transfer dyes mainly used fluorescence
spectroscopy because the photoproduced K* (or Z*) generally Figure 7. General scheme of a reversible excited-state intermolecular
emits with a large Stokes shift (6000−10000 cm−1). Advances proton transfer (ESPT) process in excited HPTS. Reprinted with
in laser spectroscopy gave details on PES’s reactions at the permission from ref 186. Copyright 2016 Royal Society of Chemistry.

M DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 8. (A) Normalized (to the maximum intensity) fluorescence spectra of HPTS (solid line), HPTS/MCM41 (dashed line), and HPTS/
MCM41 upon addition of 20 μL of water (dotted line) in dichloromethane suspensions. Schematic illustration (not in scale) of (B) HPTS/MCM41
in a dichloromethane suspension, and (C) HPTS trapped within a water drop adsorbed on the MCM41 surface. The insets of (B) and (C) show the
wavelengths corresponding to the emission intensity maxima and the fluorescence lifetimes (τ, τRO) of the formed species and of the values of proton
transfer (kPT) and recombination (krec) rate constants. Adapted with permission from ref 186. Copyright 2016 Royal Society of Chemistry.

Figure 9. Emission decays of HPTS/MCM41 in dichloromethane suspension containing (A) water and (B) deuterium oxide. Reprinted with
permission from ref 186. Copyright 2016 Royal Society of Chemistry.

irreversible.183 In neutral conditions, the proton motion occurs When HPTS in water solutions is excited, its acidity
with a kPT = 7.6 × 107 s−1 and krec = 4.7 × 1010 M−1 s−1, significantly changes (pKa = 7.7 to pKa* = 0.6).187 In neutral
generating an anion that has a fluorescence lifetime of 9 water solutions, HPTS undergoes a reversible ESPT reaction
ns.183,184 When 2-NpOH is caged in β-cyclodextrins (β-CDs), that is characterized by kPT and krec of 1 × 1010 s−1 and 6 × 109
the ESPT reaction exhibits a reversible mechanism with kPT = Å s−1, respectively.186,188 To understand the dynamics of the
2.2 × 107 s−1 and krec = 2.3 × 1010 M−1 s−1, which are slower interaction of HPTS with SBMs, we briefly discuss its behavior
than those recorded in bulk water.182,185 The observed in other hosts such as CDs and the human serum albumin
reduction of kPT and krec values is due to the encapsulating β- (HSA) protein. The ESPT reaction of HPTS within these hosts
CDs cage having a less polar and less protic character than the follows an irreversible mechanism with a rate constant (kPT)
bulk water microenvironment does, reflecting the important that is lower than that in bulk water.189−191 The change is
role that the host plays in the PT dynamics.44,46,182 In contrast, attributed to the slow solvation dynamics of the encapsulated
the fluorescence lifetime of the caged anion (τAN = 10.5 ns) is water inside or at the pore gates of these hosts (biological
longer than that in aqueous solutions due to the protection water).189−191 Within these hosts, HPTS exhibits two different
provided by the CD cavity. 44,182 When 2-NpOH is PT pathways that are characterized by times of 150 fs and 1.2
encapsulated in solid NaX zeolites, a faster ESPT reaction ns in HSA, and 140 ps and 1.4 ns in γ-CD, reflecting the
(kPT = 3.3 × 109 s−1) and a slower recombination process (krec heterogeneity of the formed complexes.190,191 By heterogeneity,
= 8.3 × 108 M−1 s−1) were reported.179 The larger value of kPT we mean that the trapped dye molecules are found in locally
observed for these hybrid complexes, in comparison with that different environments and positions within the chemical (CD)
in water solutions (kPT = 7.6 × 107 s−1), is explained in terms of and biological (HSA protein) pockets. For the HPTS:HSA
a stronger basic character of the NaX host. The formation of complexes, the shortest time is due to a strong interaction of
strong H-bonds between the silicate framework (Si−O−Si) and the dye with the host in which a direct reaction with the
trapped 2-NpOH molecules leads to a fast deprotonation of the carboxylate groups of the protein’s amino acids is suggested.
latter.179 When the trapped anion interacts with the NaX The slower process is assigned to a PT reaction with the
material, it lives 4.7 ns, which is significantly shorter than in biological encapsulated water molecules, which have slower
water solutions (9 ns), reflecting an enhanced intersystem dynamics.190,191 These results once again show the importance
crossing in the zeolite nanocavities.179 of the host’s nature on the photobehavior of trapped H-bonded
Another well-known photoacid molecule, 8-hydroxypyrene- aromatic molecules.
1,3,6-trisulfonate (pyranine, HPTS, Scheme 3), was inves- However, HPTS does not undergo PT when it interacts with
tigated in the presence of mesoporous MCM41 materials.186 the surface of MCM41 materials in dichloromethane
N DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

suspensions, and thus, the emissions of the protonated


monomer and adsorbed aggregates were recorded (Figure
8).186 The emission spectrum is centered at 430 nm, and the
fluorescence lifetimes of these species are 350 ps and 2.5 ns,
respectively. Adding water (1%) to these dichloromethane
suspensions leads to a reversible ESPT reaction and a strong
emission of the anion.186 In this case, the PT dynamics are
slower than those observed in bulk water, with PT and
recombination rate constants kPT = 5.4 × 109 s−1 and krec = 2.2
× 109 Å s−1, respectively.186 The decrease in the values of the
rate constants is explained in terms of a decrease in the HPTS
photoacidity due to a partial loss of the negative electronic
charge of its SO3− groups that are H-bonded to the
framework.186 Furthermore, the immobility of the HPTS Figure 10. Proposed diagram of proton transfer and relaxation of the
anions on the silica frameworks leads to a slower recombination prototropic species of excited-state 6-HQ in microporous catalytic
process.186 When normal water is replaced by deuterated one, faujasite zeolites. Adapted with permission from ref 199. Copyright
the PT and recombination processes of these complexes are 2010 Wiley-VCH.
even slower, with kPT = 2.7 × 109 s−1 and krec = 1.7 × 109 Å s−1
(Figure 9). The study reported that the kinetic isotope effects
(KIEs) are 2 for (kHPT/kDPT) and 1.3 for (kHrec/kDrec), being intensities of AN* and Z* bands depend on the properties of
larger for the PT process than the normal one in bulk water the zeolite’s cage: AN* emission is predominant in NaX, while
(1.4).186 Thus, the H-bonding interactions between HPTS and that of Z* in NaY.199 The authors explained this difference in
the MCM41 framework and the restricted diffusion space terms of the zeolite basicity. Since NaX is more basic than NaY,
available to the adsorbed HPTS result in slower dynamics.186 the AN* population is larger in its cages. The ESPT times and
However, a different behavior was observed when HPTS the values of the fluorescence lifetimes of the different formed
interacts with other SBMs such as alumino-silica surfaces and species of 6-HQ in these zeolites (kN−1, kAN−1, and kT−1 for the
hydrated porous silicon (10 nm pore size).188,192 For both neutral, anion, and zwitterion forms, respectively) are shown in
complexes, the ESPT mechanism is similar to that observed for Table 1.199
HPTS in bulk water. This similarity is because the HPTS Clearly, the zeolite confinement decelerates the forward
molecules are covalently linked to the alumino-silica surfaces reactions (from 20 to 70 ps for k1−1 and 43 to 400 ps for k2−1)
and retain their photoacid character, while in the hydrated and accelerates the backward ones (from 380 to 140 ps for
porous silicon, the large pore size of the host does not affect the k−1−1 and 1 ns to 750 ps for k−2−1),199 when compared with the
ESPT dynamics.188,192 Clearly, the different behaviors of HPTS rate constants in water solutions.202,203 These photophysical
when interacting with these materials reflect how the molecular changes reflect that the zeolite confines the contact ion pairs to
environment and the restriction of the host affect the ESPT be closer than they are in bulk water, thus enhancing the
reaction of the PT dye and its related dynamics. forward reactions.199 Note that within CDs, the forward
The hydroxyquinoline family, having amphoteric character, reactions (380 and 360 ps) are even slower than in the zeolite
has been intensively studied due to the possibility of assisted complexes.202 Confined water in the CD cavities leads to slower
ESPT reactions with protic media (water, alcohols, acetic acid, molecular dynamics.44 However, within zeolite cavities, where
or ammonia, among others).138,139,193−198 As a result, there is no confined solvent (or its population is very weak),
hydroxyquinolines were selected as good candidates to study the dynamics are only affected by the specific and nonspecific
the catalytic properties of SBMs as single and double PT interactions with the acidic and basic sites of the framework.
reactions can occur in the pores of these materials.143,144,199−201 These examples nicely show the relevance of the medium
In a neutral aqueous solution, 6-hydroxyquinoline (6-HQ, (solution, suspension, or solid) in exploring the photodynamics
Scheme 3) exhibits an ESPT reaction in a stepwise fashion: an of these confined systems. On the other hand, the kinetic
initial proton release from the OH group of 6-HQ that constants of the PT and recombination processes of 6-HQ
produces the anion (AN) in ∼20 ps (E*→ AN*) followed by a within NaX and NaY zeolites are also affected by the acid/base
PT to the nitrogen atom of AN* that forms the final tautomer properties of these hosts (Table 1).199 The enol deprotonation
(AN*→ T*) in ∼45 ps (Figure 10).202,203 process (k1−1) is slightly faster in NaX, a zeolite with a high
When caged in β-CD, 6-HQ shows a similar mechanism.202 basicity, than in NaY (67 vs 71 ps, respectively), while the
The forward reactions (k1 and k2), with time constants of ∼380 reprotonation (k−1−1) is slower (180 vs 120 ps, respectively).
ps and ∼360 ps, are 19 and 8.4 times slower than those in Similarly, the anion protonation (k2−1) is slower in NaX than in
water, respectively.202 However, the reverse PT events (k−1 and NaY, and its redeprotonation (k−2−1) is faster.199 In NaX
k−2) are slightly accelerated by CD’s encapsulation.202 When complexes, the predominant species is the anion as it is
trapped in NaX and NaY zeolites, 6-HQ exhibits a similar reflected by its large contribution (at ∼400 nm) to the time-
reversible ESPT process.199 The UV absorption spectra show resolved fluorescence spectra (Figure 11).199 Thus, the host
three bands centered at 325, 365, and 430 nm, indicating three properties play an important role in determining the pathway
possible prototropic species (E, AN, and Z, respectively) in the and the rate constants of the ESPT reactions: the OH-group
S0 state. In contrast, the emission spectra only exhibit two deprotonation processes are the most favored in a basic host
bands, which have intensity maxima at 400 and 510 nm, (NaX), while the imine protonation processes are favored in
assigned to the emission of AN* and Z*, respectively. No more acidic ones (NaY).199 These results are in agreement with
emission signal was detected from E*, suggesting that almost all the fact that the protonation of the 6-HQ anion is largely
the molecules in this form readily deprotonate. The relative favored in the presence of acetic acid.195
O DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Table 1. Kinetic Constants of Excited-State Tautomerization and Relaxation of 6-HQ Encapsulated in Zeolitesa
zeolite k1−1 (ps) k−1−1 (ps) k2−1 (ps) k−2−1 (ps) kN−1 (ps) kAN−1 (ps) kT−1 (ps)
NaX 67 180 420 710 1100 9600 15600
NaY 71 120 380 770 1450 12200 18000
a
Reprinted with permission from ref 199. Copyright 2010 Wiley-VCH.

interact with the labile protons of the E form of 7-HQ,


stabilizing the trapped AN.143 The emission spectra of these
composites are rather complex, displaying signals from 325 to
650 nm. On the basis of the excitation-dependent emission
spectra and the lifetime distributions, the authors assigned the
emission to E*, AN*, and Z* that are all caged and H-bonded
to the framework. Their fluorescence lifetimes are 0.26, 1.5, and
5.5 ns, respectively.144 Femtosecond emission studies of caged
E* revealed the deprotonation and protonation dynamics
within the MCM41 and AlMCM41 composites (Figure
12).143,144 The obtained time constants are 0.3 and 3 ps,

Figure 11. Time-resolved fluorescence spectra of 6-HQ trapped in (A)


NaX and (B) NaY after a time delay of 0 (purple), 70 (blue), 200
(green), 500 (yellow), and 2000 ps (red). The inset in (A) is a cartoon
(not to scale) of 6-HQ encapsulated in a faujasite zeolite nanocavity.
Reprinted with permission from ref 199. Copyright 2010 Wiley-VCH.

7-Hydroxyquinoline (7-HQ, Scheme 3), a well-studied


quinolone derivative, also exhibits single and stepwise double
PT reactions.143,144,200 In water solutions, the OH-deprotona-
tion and imine protonation processes occur in both S0 and S1
states and are reversible in the former.204 Upon electronic
excitation of 7-HQ (to S1), the time constants of these acid and
base processes are 40 and 160 ps, respectively.204 Several
groups have reported that the photobehavior of 7-HQ Figure 12. Femtosecond (fs)-emission transients of 7-HQ interacting
prototropic forms is sensitive to the confining environ- with MCM41 in a dichloromethane suspension. Reprinted from ref
ment.125,139,185,187,204,205 For example, when 7-HQ is caged in 144. Copyright 2010 American Chemical Society.
β-CD, its OH-deprotonation is slower (170 ps), while its imine
protonation is faster (85 ps) than those observed in water
solutions.39,204 Moreover, in these cavities, the enol population respectively, in MCM41, while in AlMCM41, both processes
is the predominant form, while in pure water, the zwitterions are slower (0.5 and 6 ps, respectively), showing the importance
dominate. This difference reflects the hydrophobic nature of of specific and nonspecific interactions with the nanohosts in
the CD interior stabilizing the enol form and the slowing down this guest’s dynamics and spectroscopy.143 Both processes are
of the double PT processes that occur at the two gates of CD significantly faster than those in water (40 and 160 ps) and CD
cage.185 Similar behavior was observed in nonaqueous protic (170 and 85 ps) solutions, which emphasizes direct PT events
solvents, such as alcohols: an H-bond bridge is formed with the in robust composites, where confinement, specific, and
enol to function together as a proton relay.139 Solvent nonspecific interactions play a major role in its photobehavior.
reorganization to reach an appropriate 7-HQ/solvent config- Because of these interactions with the host, even molecules
uration is an important factor to consider in PT reac- that exhibit ultrafast ESIPT reactions in solutions can undergo
tions.139,184,206,207 ESPT processes when they are encapsulated in micro- or
When 7-HQ is within regular MCM41 materials in mesoporous materials, making very rich their spectroscopy.
dichloromethane suspensions, its OH-deprotonation and This possibility is illustrated by the cases of 2-(2′-
imine protonation are irreversible.144,179 Contrary to the hydroxyphenyl)benzothiazole (HBT), 2-(2′-hydroxyphenyl)-
dynamics in bulk water, both the AN and Z/K forms are benzoxazole (HBO), and other derivatives when they interact
stabilized in S0 when the dye is caged in MCM41 pores, with zeolites and mesoporous materials (Scheme 3).208−214 In
showing how the hydrophilicity, structure, and confinement of solutions and upon electronic excitation, this kind of molecules
this host affect the formation and stability of these exhibits ultrafast ESIPT reactions (<50 fs) to produce K*
species.143,144 Moreover, the presence of Al atoms in tautomers, which exhibit an emission with a large Stokes
MCM41 (AlMCM41) highly stabilizes the AN form in S0.143 shift.145,146,167,215−219 Several studies have demonstrated that
The incorporation (doping) of Al atoms into the silica the ESIPT mechanism is also affected by solvent properties: in
framework introduces negatively charged sites that are able to polar and protic media, intramolecular and intermolecular H-
P DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 13. (A) Schematic representation of HBO structures in S0 in dichloromethane solution (Esyn and Eanti) and interacting with NaX/NaY
zeolites [Esyn bonded to the surface (E′) and anion populations within the supercages (AN1) and within the microchannels (AN2)]. (B)
Femtosecond-emission transients of HBO interacting with NaY and NaX in dichloromethane suspensions. Reprinted from ref 213. Copyright 2014
American Chemical Society.

Figure 14. Steady-state emission spectra of (A) HBTNH2−SiO2, (B) HBTNH2-MCM41, and (C) HBTNH2−AlMCM41 in acetonitrile
suspensions. The insets show the composites structure. Adapted from ref 209. Copyright 2010 American Chemical Society.

bonding interactions compete.215,216,220−225 Therefore, one can ∼13 Å in diameter) and 2.5 ns (AN2 trapped in nanochannels
expect that the encapsulation of these molecules in SBMs, of ∼8 Å in diameter that connect the supercages), are observed.
which are polar and hydrophilic in their regular structures, may The greater restriction of motion within the nanochannels, due
generate rich spectroscopy and dynamics of different confined to the smaller pore diameter, precludes the possibility of a
PT reactions. This richness was observed in HBT.107,111 In conformational change (twisting) in the confined anions and
alkaline water and methanol solutions, this molecule undergoes thus increases its emission lifetimes. The presence of different
ESPT in 1.6 ps and becomes an anion in 4.4 ps.212 When HBT anion populations, compared with the single anion observed in
is encapsulated in faujasites (FAUs) and mordenite framework alkaline methanol and ethylene glycol solutions (τAN = 3 ns),
inverted (MFI) zeolites, the ESPT reaction occurs in 1.5 reflects the heterogeneous distribution of HBO molecules
ps.208,212 This result demonstrates that the deprotonation resulting from different specific and nonspecific interactions
processes in water solutions and in FAU zeolites are not very with the framework.213
different, suggesting that in the composites, the encapsulation An HBT derivative, 2-[5′-N-(3-triethoxysilyl)propylurea-2′-
generates a PES that is not very different from that in water.212 hydroxyphenyl]-benzothiazole (HBTNH2, Scheme 3), was
When HBO, a molecule structurally comparable to HBT, is covalently bonded to the surfaces of silica nanoparticles and
confined within FAU zeolites, a similar photobehavior was mesoporous materials (Figure 14).209 In these complexes, an
observed.213 In alkaline methanol and ethylene glycol solutions, ESPT reaction took place between the dye and the silica
it exhibits an ultrafast ESPT reaction (<100 fs) that is followed framework.209 Steady-state experiments showed that in the
by ICT (1.2 ps) and a twisting motion (12 and 20 ps in amorphous silica nanoparticles, anions formation occurs,
methanol and ethylene glycol, respectively).213 When it is although stronger emission is observed from the K* structure
encapsulated in NaX and NaY zeolites in dichloromethane (Figure 14A).209
suspensions, it exhibits a similar ultrafast ESPT reaction (<100 However, when HBTNH2 is bonded to the MCM41 and
fs), but the subsequent ICT process is slower (2.5 ps) and the AlMCM41 frameworks, the interactions with these hosts lead
twisting motion is now canceled (Figure 13).213 In the NaX to a greater yield of anions and an increase in its emission
and NaY complexes, two anion populations, with fluorescence intensity (Figure 14, panels B and C).209 The replacement of Si
lifetimes of 800 ps (AN1 encapsulated within supercages of atoms by Al creates negative charges on the framework, which
Q DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 15. Femtosecond emission transients of (A) HBTNH2, (B) HBTNH2−AlMCM41, (C) HBTNH2-MCM41, and (D) HBTNH2−SiO2 in
acetonitrile suspensions. Reprinted from ref 209. Copyright 2010 American Chemical Society.

increases its proton affinity and results in predominant anions methylation of their frameworks, affects the spectroscopy and
formation in the AlMCM41 materials.209 For the femtosecond dynamics of the guest.
dynamics of these HBTNH 2 /SBMs complexes, the 3.1.1.2. Molecules in S0 and S1 States that Exhibit
AlMCM41’s composites also behave differently than Protonation Processes with the Framework. Some molecules
MCM41’s ones (Figure 15).209 The lack of the ultrafast can be protonated when they are encapsulated in SBMs.229−240
components in the fluorescence transients of the composites The guest forms one or more cationic species depending on the
with Al doping indicates that only the anion dynamics is available proton-acceptor groups and the acidity of the
present. These transients have a 2.5 ps component, which is encapsulating material.231−233 In these systems, intermolecular
assigned to vibrational relaxation (VR)/cooling processes, and a H-bonds may stabilize the encapsulated guest and increase its
2.5 ns lifetime assigned to the anions fluorescence (Figure fluorescence lifetime.
15B).209 Proflavine (3,6-diaminoacridine, PF, Scheme 3) is an
The HBTNH2 case and the previously discussed example of example of this kind of molecule.231−233 In aqueous solutions,
7-HQ nicely show that the replacement of few Si atoms by Al PF has a protolytic equilibrium involving mono- and dication
ones modifies the structural, chemical, and physical character- species.241 When it interacts with zeolites [Zeolite Socony
istics of the host framework (e.g., acid−base properties), which, Mobil-5 (ZMS-5) and NaY] and MCM41 materials, different
in turn, affects the guest’s photobehavior.143,209,226−228 Another photobehaviors were observed.231−233 Within the ZSM-5
zeolite, PF is present in the monocationic form (PFH+),
report on the encapsulation of 2-(4′-amino-2′-hydroxyphenyl)-
while within NaY and MCM41 materials, only its neutral form
benzothiazole (AHBT, Scheme 3) by hybrid SBMs shows how
is populated.231,232 The emission decays of these complexes are
the substitution of OH groups by methyl (-Me) ones in the
biexponential, except the case of ZSM-5, whose decay fit
silica framework modifies the photobehavior of the formed required a triexponential function.232 In the studied complexes,
composites.211 The amount of Me doping was changed from two fluorescence lifetimes (∼2 and ∼5 ns) were assigned to PF
1.4 to 9 mmol/g. Steady-state measurements showed that for molecules that have weaker and stronger interactions with the
the OH materials, the AHBT open-enol form is stabilized in host, respectively.232 The ZSM-5 zeolite (MFI structure) is the
both S0 and S1 states as a result of intermolecular H-bonding most acidic material among the zeolites used, and it is
with the host. However, for frameworks with Me-groups, the characterized by a high content of protons in its
formation of the K* tautomer is favored, and it is the main structure.4,242,243 Thus, the additional 1.3 ps component in
caged species for the sample having the highest used methyl the emission decays of the PF/ZSM-5 complexes was assigned
content (9 mmol/g).211 The presence of Me-groups in the to a protonation of PF from the framework.232 A comparable
silica-based framework increases its hydrophobic character and behavior was observed in thionine (Th + , Scheme 3)
thus decreases its capacity to H-bond with the dye. In this encapsulated in these materials.234 The fluorescence decay of
manner, the closed E conformer of the dye, which results in the Th+ in an aqueous solution shows monoexponential behavior
K* species upon excitation, is stabilized. These results with a time constant of 0.34 ns.244 However, when Th+ is
demonstrate how the nature of the acid−base properties of trapped in ZSM-5 and MCM41 materials, it exhibits a
these SBM host, changed by either metal doping or biexponential decay with time constants of 0.33 and 1.77 ns
R DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

for ZSM-5 complexes and 0.16 and 0.31 ns for MCM41.234 The nonradiative deactivation of the excited-state. When Sudan I is
similar 0.33 and 0.31 ns time constants are attributed to Th+ encapsulated in NaX, NaY, and MCM41 materials, both trans-
molecules that are adsorbed on the materials’ surface. In these and cis-proton-transferred isomers emit. The emission lifetimes
composites, Th+ molecules have little movement restriction and of confined trans- (20−35 ps) and cis-HYZ (113−184 ps)
strongly interact with the water molecules of the medium.234 structures are increased by 2 orders of magnitude, which is
The 1.77 and 0.16 ns times are assigned to encapsulated mainly due to the spatial and electronic confinement in these
monomer and dimer species, respectively. A femtosecond study hosts.260,261 The trans-HYZ lifetime (21 ps) in MCM41, which
of Th+ in ZSM-5 revealed that the encapsulated monomer has a larger pore size (25 vs 13 Å for zeolites), is shorter than in
species undergoes an interconversion between two protonated NaY (34 ps) (Figure 16).261 Similar behavior was observed for
forms in 2.6 ps. However, within MCM41, the protonation the cis-HYZ form, whose fluorescence lifetime was longer in
process that was described above for the PF molecule does not NaY materials (184 ps) than in MCM41 complexes (113
occur.234 ps).261
Other molecules, such as ketones and benzophenone, have
been used as probes to characterize new solid sup-
ports.236−240,245−248 Benzophenone (BZP, Scheme 3) is widely
used to investigate the acidic character of the environ-
ment.247,248 As in the previous composites, BZP that is
encapsulated in materials that contain Brønsted or Lewis acid
sites [ZSM-5 and aluminophosphate (AlPO4) microporous
solid] exhibits protonation processes.238,239 In BZP/AlPO4
complexes, the luminescence spectra show two bands (440
and 470 nm) that are assigned to the H-bonded and protonated
BZP species, respectively. The behavior was compared with
that of BZP encapsulated in ZSM-5, a host that is less
hydrophilic and polar than AlPO4.238,239 When Na+ cations are
introduced into the ZSM-5 framework, Lewis acid sites are
favored. In sodium-doped ZSM-5 materials, in addition to the Figure 16. Fluorescence decays of Sudan I in the different meso- and
H-bonded and protonated species (formed from the interaction microporous materials gated at 580 nm. Reprinted with permission
with the Brønsted acid sites of the host), the formation of from ref 261. Copyright 2009 Elsevier B.V.
radical species was reported.238 The relative amounts of H-
bonded, protonated, and radical species depend on the number On the other hand, fs-studies indicated that the photo-
and nature of the acidic sites in each material, making BZP a isomerization process in these cages occurs by an inversion
sensor that provides relevant information about the acidic mechanism, as the emission transients have similar dynamics
characteristics of the host, which is of great importance in the that are independent of the host size (Figure 17, panels A and
catalysis field.236−240 B). For the fs-emission transients of these complexes, two time
3.1.2. Intramolecular Proton-Transfer Reactions. As constants, 240 fs (IVR) and 1.70 ps (isomerization of the trans-
said at the beginning of this section, the ESIPT reaction occurs HYZ structure along N−N and/or C−N bonds), were
in organic bifunctional molecules that have both proton donor reported. The presence of solvent (water and dichloromethane)
and acceptor groups in close proximity.98,101 The dynamics of in the nanochannels and different metal cations (Na+, Li+, and
the ESIPT reaction may depend on the environmental Mg2+) in the framework affect the isomerization time (Figure
properties such as its polarity, viscosity, and ability to form 17C).260 In LiX and MgX zeolites, the Li+ and Mg2+ cations
H-bonds in the same manner as the ESPT reac- slow the trans-cis isomerization process (2.8 and 2.3 ps,
tion.117,120,161,215,216,220,221,249−254 respectively),260 while Mg2+ cations (MgX zeolite) presence has
Understanding the ESIPT reaction and the possible a smaller effect than the Li+ one.260 These results show that the
subsequent processes such as rotation or isomerization is nature of the metal cation (size and charge) is paramount to the
paramount to deciphering fundamental atomic motions but electronic charge density inside the host and affect the
also for designing new materials for photoswitches, sensors, photobehavior of the trapped dye.
energy storage and nanomedicine devices, to cite a few Salicylaldehyde azine (SAA, Scheme 3), another diazo-
applications.43,255−258 For example, the encapsulation of derivative, was studied in FAU zeolites.263 In dichloromethane
photochromic aromatic molecules, which undergo ESIPT and solutions, SAA exhibits an ultrafast ESIPT reaction (80 fs) that
subsequent isomerization, is of great interest to the leads to a K*-type tautomer having an emission lifetime of 55
fundamental knowledge and applications.259−273 Sudan I (1- ps.263 The Z form of SAA is stabilized in NaX zeolites in
phenylazo-2-naphthol, Scheme 3), which belongs to the group dichloromethane in both the S0 and S1 states to the point that
of diazo-conjugate dyes, has been studied in FAUs (NaX, LiX, no ESIPT reaction occurs.263 In the SAA/NaX complexes, the
MgX, and NaY) and MCM41 in several solvent suspen- Z* fluorescence decays exhibit a triexponential behavior (0.18,
sions.260,261 In solution, it exhibits an ESIPT reaction (<50 fs) 0.9, and 2.8 ns) that is slower than SAA’s fluorescence in
that in S1, produces a trans-ketohydrazone (trans-HYZ) solution, which is mostly because the twisting motion of the
structure, which subsequently isomerizes (by rotation, inver- confined structures is hindered.263 The fluorescence lifetime
sion, or some mixed mechanism) in an ultrafast fashion to a cis- values significantly depend on the SAA concentration in these
ketohydrazone (cis-HYZ) structure that has a fluorescence complexes (Figure 18).263
lifetime shorter than 2 ps (Scheme 3).259 The shorter lifetime The presence of three fluorescence lifetimes and their
of cis-HYZ is due to a twisting motion of the two aromatic parts dependence on the dye concentration were explained by either
(i.e., along −NN− or N−NH− bonds) that enhances the caged SAA structures having different locations or by a self-
S DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 17. (A and B) Femtosecond emission transients of Sudan I in NaX, NaY, and MCM41 hosts in n-hexane suspensions. (C) Femtosecond
emission transients of Sudan I in MX (M = Na+, Mg2+, and Li+) in neutral water gated at 620 nm. Reprinted from ref 260. Copyright 2009 American
Chemical Society.

Figure 18. Effect of the used initial SAA concentration (in the
synthesis of the SAA/NaX composite) on the fluorescence decays
observed at 480 nm. Reprinted from ref 263. Copyright 2010
American Chemical Society.
Figure 19. Emission decays of dichloromethane solutions containing
HBA-4NP interacting with NaY using (1) concentrated (2 × 10−3 M)
quenching process. To analyze the quenching effect, the and (2) diluted (2 × 10−7 M) initial HBA-4NP solutions, observing at
fluorescence decays were fitted to a stretched-exponential 500 nm. Reprinted from ref 264. Copyright 2014 American Chemical
function.88,90−92 This function is also used in modeling EnT Society.
processes presented in section 2.2.3. Upon decreasing the SAA
concentration in the composites, the characteristic time of the due to a hindering of the nonradiative pathways.264,265 On the
decay, τs, changes from 0.19 to 1.67 ns, while the heterogeneity other hand, the relative populations and emission lifetimes of
parameter α increases from 0.46 to 0.61. The shortening in the H- and J-aggregates (∼100 and ∼600 ps, respectively) were
emission lifetime indicates possible interaction and EnT found to depend on several factors such as the loading
between neighboring caged SAA molecules.263 Therefore, the efficiency, the initial concentration of the dye used to make the
multiexponential behavior of the Z* tautomer can result from complexes, cavity (or pore) size, and the Brunauer−Emmett−
the formation of different populations inside the NaX cavity Teller (BET) surface area of the materials.264,265 For
and/or from different interactions between caged SAA microporous materials, HBA-4NP exhibits larger aggregates
molecules. The confinement effect of NaX on SAA is also formation in NaY than in NaX due to the higher loading of the
reflected at the femtosecond scale: the ultrafast IVR (∼300 fs), former (90 vs 20%, respectively), which is related to the lower
VR/cooling (∼5 ps), and twisting motion (∼10 ps) are slower content of sodium cations and aluminum atoms in the zeolite
than those in dichloromethane solutions (∼100 fs, ∼1 ps, and framework.264 For the mesoporous materials, whose pore size
∼5 ps, respectively).263 The salicylideneaniline derivative (E)-2- (25 Å) is larger than that of zeolites (13 Å), higher loading (85
(2′-hydroxybenzyliden)amino-4-nitrophenol (HBA-4NP, and 95% for MCM41 and AlMCM41, respectively), and thus
Scheme 3) behaves similarly to SAA in the presence of FAU more aggregates formation was observed.265 However, within
zeolites, NaX, NaY, and mesoporous materials such as MCM41 SBA15 host (pore size = 60 Å), there is less aggregates
and Santa Barbara amorphous (SBA15) type material.264−267 In formation (40%) due to its weak loading ability, which is
dichloromethane solutions, HBA-4NP undergoes an ultrafast related to its BET surface area (780 m2/g for SBA15 vs 1000
ESIPT reaction (∼70 fs) that leads to a K*-type tautomer with m2/g for MCM41).265 Figure 20 illustrates the different loading
a short lifetime (14 ps).266 The photoproduced K* undergoes a of MCM41 and SBA15 materials. For complexes having a low
rotational process (∼4 ps) toward nonemitting structures.266 loading efficiency (NaX and SBA15 hosts), the caged
When HBA-4NP interacts with SBMs, the shape and position monomers are the predominant guest.264,265 In contrast, for
of the features in the absorption and emission spectra, as well as complexes resulting in a high loading efficiency (NaY, MCM41,
the fluorescence lifetimes, show a large dependence on the dye and AlMCM41 hosts), H- and J-aggregates are more commonly
concentration (Figure 19). formed, with the latter being predominant in the presence of Al
The analysis of the data suggests the presence of monomeric atoms.264,265 This result demonstrates that depending on the
or H- and J-aggregate caged E structures that undergo ESIPT structural and topological properties of the SBMs, different
reactions.264−267 The fluorescence lifetime of the trapped K* encapsulated species of HBA-4NP might be formed.265 The fs-
monomers (∼6 and ∼2 ns for zeolites and mesoporous hosts, photobehavior of these complexes also depends on the
respectively) is longer than that in dichloromethane (14 ps) host.266,267 For the NaX and SBA15 composites, the ultrafast
T DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 20. (A) Schematic representation (not in scale) of the energetic diagram at ground (S0) and electronically first excited (S1) states of
monomer (M) and H- and J-aggregates of HBA-4NP interacting with mesoporous materials in dichloromethane suspensions. For simplicity, the (S0)
and (S1) energy of the three caged structures are put at the same level. (B) Cartoon illustrating the caged structures (M and H- and J-aggregates)
within MCM41/AlMCM41 and SBA15 materials. The fluorescence lifetime for each specie is indicated in the scheme. Reprinted with permission
from ref 265. Copyright 2016 Elsevier B.V.

dynamics are dictated by the encapsulated monomers, similar anion form of 3-HF.269 However, within polar and protic
to those observed in viscous triacetin solvent, while for NaY, zeolites such as β- and HZSM-5, the encapsulated 3-HF does
MCM41, and AlMCM41, it is rather due to the aggregates, not transfer protons and only the fluorescence of the
which photodynamics is slower than in solution.266,267 For encapsulated N* form was observed.271,272 However, in aprotic
these hybrid composites, a larger amount of HBA-4NP caged and polar materials such as NaZSM-5 zeolites, both N* and T*
inside the host leads to guest−guest interactions that, along structures are present in the excited state.271 Time-resolved
with the restriction of twisting motion, slow the guest’s global diffuse reflectance spectroscopy showed that the fluorescence
ultrafast events such as IVR and VR/cooling processes. In the lifetimes of N* and T* species are very long, with values of 165
mesoporous MCM41 and AlMCM41 complexes, the dynamics and 746 μs, respectively.271 However, when 3-HF interacts with
are slower than those in microporous NaY due to the wider polar and aprotic materials such as silicalite-1, it does not
space and higher content of OH-groups in the former, which interact with the channel walls, and T* is the only observed
result in a larger number of specific interactions.267 species (τT = 500 μs).271
Hydroxyflavones are another family of molecules that The number of reports discussed in the PT section shows the
undergo ESIPT reactions. Among them, 3-hydroxyflavone (3- importance of the nature of the SBMs when studying the
HF, Scheme 3) has drawn particular attention since its spectroscopy and dynamics of a PT dye. The interaction
spectroscopy depends on its surroundings, and it may be between both entities can be specific or nonspecific, leading to
used in several fields of science and technology.97,105,274 While caged monomers, anions, cations, keto forms, zwitterions, and
in aprotic and apolar solutions, it undergoes an ESIPT reaction aggregates. The diversity of these composites is reflected in
in the photoexcited neutral (N*) form to produce the tautomer their photodynamics and spectroscopy, which greatly help the
(T*), and in protic and polar solutions, the formation of T* is understanding of many host−guest systems in both the ground
inhibited by the formation of intermolecular H-bonds with the and excited states. In general, the observed caged photo-
solvent.274,275 Several studies have reported on 3-HF in the dynamics is longer than that in solution.
presence of caging media as micelles,268 CDs,269,270 sol−gel
3.2. Electron-Transfer Reactions
glasses,276,277 and liposome membranes.278,279 Upon excitation
of 3-HF in micelles, an equilibrium is reached between the N* Photoinduced electron-transfer (ET) reactions are one of the
and T* species reflecting a reversible ESIPT reaction.268 In fundamental processes in physics, chemistry, and biology.280,281
these hosts, 3-HF exhibits the same behavior as in the H- They have been extensively studied both theoretically172,282−284
bonding solvents, such as methanol and water, indicating strong and experimentally.2,35,37,41,42,285 The standard theoretical
specific interactions between the dye and the micelle.268 In γ- model that describes ET reactions was developed by Marcus
CD, a more hydrophobic medium, an ESPT reaction produces and proved a relation among the difference in the Gibb’s free
an excited anion (τAN = 2.14 ns).269 The deprotonation occurs energy (ΔG) between an electron donor (D) and acceptor (A),
in the cavity because of specific interactions (H-bonding the curvature of the PES, and the reorganization energy.283,286
network) between the dye, CD, and adsorbed water According to Marcus theory, the energies of the electron D and
molecules.269 The hydrophobic effect imposed by the inside A states are adjusted to match by energy fluctuations that are
γ-CD cavity enhances and stabilizes the photoproduced caged caused by the surrounding media. ET can occur between D and
U DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 4. Molecular Structures of the Molecules Interacting with SBMs and Discussed in the Electron Transfer Section 3.2

A in separate molecules (intermolecular) or between distinct D discuss the key properties of these hosts that govern the rate
and A regions within the same molecule (intramolecular). In constants and subsequent processes.
such inter- and intramolecular processes, the reaction products 3.2.1. Intramolecular Charge-Transfer Reactions.
are usually called the charge-separated (CS) and ICT states, 3.2.1.1. Dependence of ICT Reactions and Solvation
respectively.2,41,287,288 The stability of both states and the value Dynamics upon Encapsulation in Silica-Based Materials.
of the ET rate constant also depend on the properties of the Molecules with electron D and A groups in their structures
medium that surrounds the molecular system. Several reviews exhibit ICT reactions upon excitation. These reactions
have been published on theoretical and experimental studies on subsequently induce changes in the electron distributions,
the photoinduced dynamics of electrons (charge) transfer dipole moments, and possible conformations at the excited
within chemical and biological sys- state.36,41,287,288,291 The ICT rate constant in the molecular
tems.2,4,7,27,32,34,37,41,172,173,281,287,289,290 The most recent review system under consideration depends on the polarity of the
gives a detailed analysis of these events in solutions and helps surrounding, and the specific and nonspecific interactions with
the reader to understand the state-of-the-art in this field.32 the medium.292−295 As we are concerned here with the
Encapsulation by hosts (such as micelles, CDs, proteins, and confinement effects of SBMs, a change in the local environ-
other molecular pockets) may change reaction paths and the ment, such as polarity, H-bonding ability or cavity/pore size is
related products. Therefore, in this section, we will focus on the expected to affect the ICT reaction dynamics and products. In
effects of SBMs (as hosts) on reaction dynamics of ET and basic media, HBO (Scheme 3) undergoes an intermolecular
V DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 21. (A) Schematic representation of coumarin dyes interacting with silica−surfactant complexes. Reprinted from ref 312. Copyright 2006
American Chemical Society. (B) Schematic illustration of the mesostructured silica nanochannels containing alumina membrane to encapsulated
C153. Reprinted from ref 316. Copyright 2008 American Chemical Society.

ESPT reaction with the host to produce an anion by the surface also influence the electronic distribution in the dye, thus
deprotonation of its OH group (section 3.1).213 The presence affecting the ICT time.186 The above reports on HBO and
of the negative charge in this structure causes an ICT reaction HPTS demonstrate the influence of the interaction with SBMs
which time depends on the surrounding medium.213 In alkaline on ICT processes. The deceleration of the ultrafast processes
methanol solutions, where the HBO anion (AN) already exists changes the overall behavior of encapsulated guests, which
in S0, the photoinduced ICT process occurs in 1.2 ps. When could affect the subsequent reactions and formed photo-
HBO is trapped within NaX and NaY zeolites, it forms an products.172,173,186,190,191,298,299,304
encapsulated AN.213 Upon excitation, the confined AN* Molecules having electron D and A groups are often very
undergoes an ICT in 2.5 ps.213 The reaction time in zeolites good probes to study solvation dynamics because their dipole
is slower than in methanol due to the higher polarity of the moments significantly change upon excitation. Therefore, such
local zeolite environment and the electrostatic interactions of molecules were also employed to study solvent reorganization
AN* with the aluminosilicate framework and sodium cations in dynamics after interaction with SBMs by time-resolved
its cages.213 HPTS (Scheme 3), which has been discussed in the fluorescence methods. In bulk solvents, the solvation process
previous section, is a well-known photoacid in the field of is usually described as consisting of a sub-100 fs Gaussian
photoinduced PT. It has been studied in solutions and in inertial component and single- or biexponential longer decay
chemical (CDs and micelles) and biological hosts.186,188−191 In components of 1−10 ps duration.305 The theoretical works
aqueous solutions, in addition to an ESPT reaction, the ultrafast about solvation dynamics in nanoconfining silica-based environ-
dynamics of HPTS involves two steps of 300 fs and 3 ps, which ments were commented in section 2.1, while here, we will
are assigned to solvation of the locally excited (LE) state and comment on and discuss the experimental findings.
ICT processes, respectively.39,174,176,296,297 When HPTS is The studies before 2005 of solvation in silica materials were
encapsulated in γ-CD, these processes slow down, with a LE limited to dyes interacting with sol−gel glasses.306−308 For
relaxation and ICT times of 800 fs and 8 ps, respectively. example, this type of material was used to explore the solvation
Trapped in HSA protein and ionic liquid mixed micelles, similar dynamics of Coumarin 480 (C480), Nile Blue (NB), and C153
behaviors were observed; and ICT dynamics slowed down to (Schemes 3 and 4).306−308 Generally, the solvation dynamics in
10−200 ps.190,191,298,299 When HPTS interacts with dried confined media (10−50 Å pores diameter) showed non-
MCM41 materials, it does not exhibit any proton- or charge- exponential behavior of the emission decays, similar to that in
transfer processes.186 Aggregates formation was observed in the bulk, but all time components were slowed down. However,
formed composites, characterized by a broad emission band the time constants of the solvation dynamics in these confining
and a multiexponential behavior of the emission decays (120 media were found to be 100−200 ps,306 which is still
ps, 600 ps, and 2.2 ns).186 However, the presence of water (1%) approximately one order of magnitude faster than in CD
in HPTS/MCM41 samples decelerates both processes. Under (∼1200 ps)309,310 and micelles (∼2400 ps).311
such wet conditions, the LE relaxation and ICT dynamics As such, much slower solvation dynamics was observed in a
exhibit times of 0.8 and 5 ps, respectively.186 This retardation is mesostructured silica−surfactant nanocomposite.312 Three
due to (i) the nature of adsorbed water molecules on the coumarin dyes were studied: two of them were located in the
MCM41 surface, (ii) the electrostatic interactions of the dye vicinity of the ionic interface [coumarin 343 (C343) and
with the MCM41 surface, and (iii) the decrease of freedom of C480], while the third one (propylamide coumarin 343
adsorbed HPTS and water molecules.186 Several studies have [PAC343], Scheme 4) was found in the hydrophobic interior
shown that adsorbed and/or encapsulated water molecules of the micelles (Figure 21A). The longest solvation time
have different dynamics than those observed in bulk water constant is 7.5 ns for C480 and 7.6 ns for C343, while for
solutions.46,47,300,301 The molecular restriction to move on the PAC343 (with a smaller overall Stokes shift) it is significantly
surface affects the polarity and dielectric constant of the solvent, shorter: 4.8 ns. The different behavior for PAC343 were
resulting in a deceleration of their dynamics.44,46,47,300−303 The explained on the basis of a smaller contribution of the water−
origins of this process are also discussed in the theoretical ion interaction at the ionic interface and fewer water molecules
section of this paper (section 2.2.1). On the other hand, the trapped in the hydrophobic core of the micelles. A similar
adsorption of HPTS and its interaction with the MCM41 nanocomposite was studied using 4-aminophthalimide (4AP)-
W DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 22. (A) Schematic representation of the O−H interaction of water with charged and neutral silica surfaces. τ1 is the interfacial O−H
vibrational lifetime. Reprinted from ref 320. Copyright 2009 American Chemical Society. (B) Variation of τ1 as a function of surface charge (pH 2−
12) at different material interfaces. Reprinted from ref 322. Copyright 2017 American Chemical Society.

based molecule as a solvation probe (Scheme 4).313 The probe behavior.213 For Nile Red (NR, Scheme 4), the combined ICT
was made from a 4AP fluorophore and a 16-membered and solvation dynamics are even shortened from approximately
hydrocarbon chain. The fluorescence spectrum indicates that 1 ps in bulk dichloromethane to 200−400 fs in dried XMCM41
the 4AP moiety resides in the silica−surfactant interfacial (X = Si, Al, Ga, Zr, and Ti).227 The increase in the environment
region, with the hydrocarbon tail extended into the hydro- polarity decreases the height of the energy barrier between the
phobic core of the nanocomposite. Time-resolved studies LE and ICT states, leading to more efficient formation of the
revealed fast and slow solvation components: 290 ps and 1.75 ICT form.227 Moreover, the ICT reaction dynamics also
ns. These times are longer than those in cetyltrimethylammo- depends on the presence of different acid sites (Brønsted and
nium bromide (CTAB) micelles (100 and 380 ps)314 but Lewis character) in the MCM41 framework; these acid sites are
comparable to those in reverse micelles (340 ps and 5 ns).315 the products of doping the framework with different
The constrained water molecules near the ionic interface of the metals.227,292 For example, when NR is inside MCM41,
micelles inside the nanocomposites were found to be the AlMCM41, and GaMCM41 (containing Brønsted acid sites),
primary contributor to the dynamics, although the slow the ICT reaction occurs within 320−410 fs and a VR/cooling
dynamics exchange of the free and bound water molecules process within 3.2−3.5 ps.227 When NR is trapped in
could also have a significant influence. ZrMCM41 and TiMCM41 hosts (containing Lewis acid
C153 was used to probe solvation dynamics in alcohols of sites), its ICT state is generated in less than ∼200 fs and is
various alkyl chains confined in silica.316 Two mesoporous silica later depopulated by an electron-injection (EI, in ∼250 fs)
nanochannels [calcined-nanochannel-containing alumina mem- process to reach trap states that are formed by the d-orbitals of
brane (Cal-NAM) and trimethylsilyl-nanochannel-incorporated the doping metals.227 NR has been also studied in both β-CD
alumina membrane (TMS-NAM)] were fabricated inside the and surfactant aerosol-OT [sodium bis(2-ethylhexyl) sulfosuc-
pores of an anodic alumina membrane (Figure 21B). The cinate, AOT] reverse micelles in solutions, in which its ICT
average solvation times are approximately 10% longer in TMS- process is 2.5 and 8 times slower than that in bulk water,
NAM than in Cal-NAM. The solvent relaxation times in the respectively.317,318 For these systems, the observed slowing was
nanochannels are 2−3 times slower than in alcohols with a attributed to the polarity of the caged water molecules being
short alkyl chain (ethanol and 1-butanol). Interestingly, for lower than that of the bulk.317,318 The above studies are
alcohols with a long alkyl chain (1-hexanol and 1-decanol), no commented about in other sections of this review (sections
differences were observed with the confined solvation 3.1.2 and 3.1).
dynamics. To explain the results, the authors proposed the The ultrafast components of solvation are partially related to
formation of rigid alcohol chains for ethanol and 1-butanol by the vibrational dynamics of solvents. Sum frequency generation
means of strong H-bonding interactions and less strong Van was shown to be a useful tool to study the vibrational spectra of
der Waals interactions between alcohol molecules.316 In the water molecules at interfaces.319 Recently, experiments
contrast, the alcohols with a long alkyl chain were considered using this technique with femtosecond time-resolution were
to form a loose structure inside the silica nanochannels. used to explore the vibrational dynamics of water at different
Solvation processes were also considered for the nano- interfaces.320−322 The vibrational lifetime τ1 of the water O−H
confined molecules for which deactivation processes other than stretch at silica interfaces was found to strongly depend on the
ICT contributed to the excited-state dynamics. In such systems, pH.320 At pH > 10, most of the surface silanols are
the isolation of pure solvation dynamics is often complicated. deprotonated, leading to a strong interfacial electric field.
For Sudan I (Scheme 3), a dye showing proton-transfer and Under such conditions, τ1 values were found to be similar to
isomerization reactions at S1, trapped in zeolites and MCM41, those in bulk water, approximately 250 fs. This similarity was
the overall increase in the observed emission lifetimes (with shown to be a consequence of probing near-interface water,
respect to that in the bulk) was explained in terms of a whose contribution was screened in the presence of excess ions.
cooperative effect of confinement (restriction to twist) and In contrast, for pH = 2, the surface charge density at the silica/
solvation within the pores/channels.260 For HBO anions water interface is known to be close to zero, and at such
(Scheme 3), the solvation dynamics in bulk solvents were conditions, the vibrational lifetime increases to 570 fs (Figure
extracted. Upon confinement in NaX and NaY zeolites, 22A). The primary reason for the longer vibrational lifetime of
additional processes related to the formation of anionic forms O−H at the H2O/silica interface at the neutral surface was
at S1 complicated the interpretation of the observed fs−ps attributed to a decrease in the density of states of the low
X DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 23. Schematic representation of the photodynamics of NR interacting with Ti-AlMCM41 materials of different metals (Al and Ti) doping in
dichloromethane suspension. Reprinted with permission from ref 226. Copyright 2016 Royal Society of Chemistry.

frequency modes due to incomplete solvation of the first layer the presence of SiO2 nanoparticles in dichloromethane
of interfacial water molecules. The thermalization time constant suspensions, the emission lifetimes (∼180 ps, ∼ 800 ps, and
is similar at both low (700 fs) and high (800 fs) pHs. ∼2.5 ns) and its ultrafast dynamics (1.8 ps) are similar to those
Interestingly, a different behavior was observed at alumina observed in neat solutions, reflecting its weak interaction with
interfaces, for which the point of zero charge (PZC) occurs at this material (Figure 24A).323
intermediate pHs (6−8) than for silica.322 The vibrational
dynamics of the O−H stretch of the charged alumina/water
interface is faster than in bulk water and at a charged silica
surface (Figure 22B). The addition of excess ions was found to
have little or no effect on the time range of the vibrational
dynamics, in contrast with the behavior observed at the silica
surface.322
3.2.1.2. Electron Injection to/from Silica-Based Materials
Containing Metals. NR shows strong solvatochromism as a
result of an ICT reaction accompanied by a twisting
motion.226,227,292,293,295,317,318 This classical dye exhibits a
different behavior when interacting with tri- (Al- and
GaMCM41) and tetravalent (Ti- and ZrMCM41) metal-
doped MCM41 materials, which is related to the presence of
different types of unoccupied orbitals.227 These unoccupied d-
orbitals in Ti- and ZrMCM41 results in an EI process from the
ICT state of NR not happening in Al- and GaMCM41 hosts.
The result agrees with those of reports in which X- and Y-
zeolites (faujasites) doped with transition metals also efficiently
form trap states thanks to their d orbitals, increasing the
catalytic activities of these materials.9,227 Interestingly, when
NR was encapsulated in co-metal-doped mesoporous materials
that have two different metals in their framework (Ti-
AlMCM41), its photobehavior depends on the numbers of
the Brønsted and Lewis acid sites.226 In Ti-AlMCM41 materials Figure 24. Comparison of the fs-emission transients of TPC1
with 1% codoping of Ti and Al, an ICT process of ∼300 fs was interacting with (A) silica and MCM41, (B) Ti-nanoparticles, and
recorded.226 At this doping level, the EI process to the trap Ti-containing mesoporous materials in dichloromethane solutions.
states was not observed.226 In contrast, when the material is Reprinted from ref 323. Copyright 2011 American Chemical Society.
doped with 3% Ti and 1% Al, the ICT state forms in <200 fs
and is followed by an EI process in ∼200 fs (Figure 23).226 The However, for the TPC1/MCM41 composites, an emission
observed changes are directly related to the Ti-content in the quenching was observed in addition to the lifetimes of weakly
MCM41 framework, and the effect could be of relevance to the interacting dye molecules. The quenching is assigned to
photocatalytic activity of this material.226 intermolecular singlet−singlet annihilation between TPC1
Another electron D/A molecule that was also studied in molecules that are closely packed in the nanochannels of the
SBMs is the triphenylamine dye (TPC1, Scheme 4).323 This mesoporous materials.323 TPC1 has also been encapsulated in
molecule has been used as a light absorber in a dye-sensitized TiMCM41 (3% Ti) and TiSBA15 (SBA15 with small titanium
solar cell, and it contains electron D (triphenylamine unit) and nanoparticles within its hexagonal mesoporous structure).323
A groups (cyanoacrylic acid) in its structure.324,325 The The dynamics of the TiMCM41 and TiSBA15 complexes differ
interaction of TPC1 with MCM41 and amorphous silica from those of TCP1 interacting with SiO2 and MCM41 (Figure
nanoparticles (SiO2) was reported.323 Excited TPC1 in 24). For the TiMCM41 composites, an efficient fluorescence
dichloromethane solutions shows two lifetimes, 800 ps and quenching process occurs due to EI from the excited dye into
2.5 ns, assigned to the normal and anion forms, respectively. In the conduction band of titania domains.323 The EI reaction
Y DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 25. Comparison of fs-emission transients (gated at 700 nm) of a dichloromethane suspension containing TPC1 interacting with (A) alumina
particles of different size and at different concentrations of the dye and with Al-doped material and (B) different metal oxide nanoparticles. Inset:
illustration of TPC1 attached to metal oxide nanoparticle surface. Reprinted from ref 327. Copyright 2011 American Chemical Society.

exhibits a multiexponential behavior with times ranging from and numerous applications such as semiconductor devices,332
300 fs to 30 ps, and an average rate constant of 2.7 × 1012 s−1 photovoltaics solar cells,333,334 chemical sensors,335,336 and
when the emission is gated at 600 nm and 1.5 × 1012 s−1 when optoelectronic devices.334,337 The spectral properties of MPc
recorded at 700 nm.323 These rate constants are 2−3 times can be modified by encapsulation or by a covalent linking
larger than those in the TiSBA15 materials, indicating a within nanochannels, resulting in interesting composites for
stronger coupling between the dye and titania domains in the solar energy conversion. Palladium phthalocyanine (PdPc) has
TiMCM41 hybrid material.323 This behavior is comparable to been studied in the presence of MCM41 materials.338 The
the case in which TPC1 interacts with the classical titania P25 samples were prepared either by diffusion of the dye in
(Figure 24B).323 This phenomenon led to the study of the dichloromethane suspensions into the mesoporous material
interaction of TPC1 with TiMCM41 in complete dye- (PdPc/MCM41) and/or covalent bonding (MO-PdPc) to its
sensitized solar cells.326 In such cells, the EI process was framework.338 The absorption and emission spectra of PdPc/
found to be approximately 3 times slower than in a typical solar MCM41 are similar to those of PdPc complexes. However,
cell configuration with P25. Unlike NR, TPC1 also exhibits an when PdPc is covalently bonded to MCM41, both absorption
ET reaction in AlMCM41 materials.327 This reaction occurs in and emission spectra are red-shifted, reflecting a change in its
1.5 ps, and it reflects the interaction between the dye and the photobehavior. This response is also manifested in the
Brønsted acid sites of the AlMCM41 framework that are fluorescence lifetime of the composites in dichloromethane
generated by the Al atoms.327 TPC1 has an even slower ET suspensions. For PdPc and PdPc/MCM41 samples, lifetimes of
reaction (150 ps) when it interacts with Al2O3 nano- and 20 ps and 2.4 ns were observed; while for MO-PdPc samples,
microparticles (Figure 25A). The difference is explained in decaying components of 20 ps, 5.8 ns, and 1.4 ns were
terms of a change in Al-orbital configurations in MCM41 and recorded.338 The 20 ps component is due to intersystem
Al2O3 nanoparticles. Replacing the Si atoms with isomorphic Al crossing (ISC) to the triplet state, which is enhanced by Pd
ones introduces a Brønsted acid site that in turn increases the atoms and is common in MPc photochemistry. The second
electron-accepting ability of the MCM41 frameworks and component is assigned to emission from a photoproduct that is
hence accelerates the ET reaction.327 In other nanoparticles formed by a photoinduced metal EI process making this
that are of interest of photonics [e.g., solar cells, light-emitting generated species living longer when the PdPc is covalently
diodes (LEDs), and photocatalysis] such as ZnO and ZrO2, the linked to the host (5.8 for MO-PdPc vs 2.4 ns for PdPc/
EI process is faster (4 and 11 ps, respectively) than that in the MCM41). The 1.4 ns component, which is not observed in
Al2O3 ones (Figure 25B).327 The involved kinetics is directly PdPc and PdPc/MCM41 complexes, is suggested to be due to a
related to the conduction band (CB) energy of these oxide different population of PdPc molecules having established
nanoparticles.327 specific interactions with the mesoporous framework. The
This phenomenon was also observed when a zinc-porphyrin results of the femtoseconds study show evidence for two
(YD2-o-C8, Scheme 4) was trapped in layers of titania and processes in the ultrafast regime: a 170−500 fs component
alumina NPs.328,329 In the presence of titania, YD2-o-C8 assigned to IVR, internal conversion, and photoinduced loss of
exhibits an ultrafast EI process through a CS state that leads to Pd ions and a 1.5−4.4 ps component due to VR/cooling
efficient fluorescence quenching. However, this process does processes.338 Therefore, the study shows that covalently
not occur in alumina because its CB lies approximately 4 eV bonding PdPc to the host redshifts the absorption and emission
above that of titania.329−331 Therefore, for YD2-o-C8/Al2O3 spectra, affects its dynamics, giving a metal-ejected photo-
nanoparticles, only an ICT state whose formation is faster in product and forming a population based on its specific
the presence of nanoparticles (500 fs) than in acetonitrile interactions with the framework.
solution (2 ps) was observed.328 These results provide Another aspect of the overall ET reactions to consider in any
information about the interaction of this dye with the system is the back ET that may happen (or electron
semiconductor NPs, which are of interest to solar cell recombination). In accordance with Marcus theory, solvent
applications (see also section 5.3). reorganization enhances back ET reactions.280 The time
Metal phthalocyanines (MPc) are an attractive class of constants of this back reaction in solutions are from
molecules that are known for their strong absorption in the femtoseconds to hours or days depending on the kind of
visible region, easy synthesis, flexibility in their molecular frame, system under study and surrounding conditions. The
Z DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 26. Cartoon illustrating proposed photoinduced processes of the radical cation of t-St trapped in zeolites cages showing: (A) charge
separation and direct charge recombination and (B) charge separation, electron−hole recombination, and hole transfer to the material. Reprinted
from ref 341. Copyright 2012 American Chemical Society.

encapsulation of dyes in SBMs decreases the probability of this lifetime also decreases with the increasing size of the alkali
reaction.28,91,326,339−351 Trapping an electron D−A molecule in metal cation (M+). For example, in M-MFI composites (M =
SBMs retards the back ET, effectively enabling the photo- Na+, K+, Rb+, and Cs+), the lifetime decreases in the following
generation of long-lived CS species. For example, when the 9- order: Na+ (22 min) > K+ (20 min) > Rb+ (16 min) and Cs+
mesityl-10-methylacridinium ion (Acr+-Mes, Scheme 2) is (12 min). However, interacting with zeolite beta shows a
incorporated into AlMCM41 materials (Acr + -Mes/ different behavior.340 In this material, the formation of the t-
AlMCM41), nanosecond time-resolved absorption and electron St•+ radical cation is faster than in previously mentioned
spin resonance (EPR) experiments on the solid-state samples zeolites because the caged t-St has greater freedom of motion.
indicated the formation of an extremely long-lived ICT state The lifetime of the CS state is tens of microseconds.340
(10 s, Acr • -Mes •+ ). 342,352 The femtoseconds transient Therefore, the available space inside zeolite hosts plays an
absorption spectra of Acr+-Mes/AlMCM41 composites reveal important role in stabilizing the CS states. This trend was also
an ICT reaction with a rate constant of 3.4 × 1011 s−1, which is observed for t-St caged in MOR zeolites doped with TiO2
slightly lower than that observed for Acr+-Mes in acetonitrile nanoclusters.358 Spontaneous ionization generates a CT
solutions (5.8 × 1011 s−1).342 Similar behavior was observed in complex in MOR zeolites without titania, but in the presence
1,6-diphenyl-1,3,5-hexatriene (DPH, Scheme 4) molecules that of TiO2, the t-St radical (70 min) is more stable. The reduction
are encapsulated in sodium-exchanged ZSM-5 zeolites, having of the internal microporous volume that is caused by the
the formula Nan(SiO2)96‑n(AlO2)n (n = 60.6 or 3.4), which are presence of titanium slows the charge recombination process by
denoted NanZSM-5.353 DPH is known for its fluorescence the confinement effect.358 In general, the introduction of ET
properties and particularly for its use as a probe in biological molecules into the restricted spaces of zeolites and mesoporous
membrane studies.354−356 Upon adsorption of DPH in materials slows back ET reactions, which is advantageous for
Na n ZSM-5 zeolites and photoionization, its CS state the development of materials for photonic applications.
(DPH•+/NanZSM5•−) is stabilized and exhibits a long lifetime 3.2.2. Intermolecular Electron-Transfer Reactions.
(minutes). This lifetime depends on the temperature, and the 3.2.2.1. Encapsulated Electron Donor and Acceptor Molec-
content and distribution of the Al atoms in the zeolite ular Systems. The study of ET reactions has been extended to
structure.353 For DHP/Na6.6ZSM-5 composites, the CS state’s bimolecular systems, in which two different molecules, one an
lifetime ranges from 3850 min at 293 K to 8 min at 353 K, electron donor and the other as an acceptor, are encapsulated
while for DHP/Na3.4ZSM-5, the lifetime changes from 65 min in the material. Electron donor t-anethole (t-An, Scheme 4) and
at 293 K to 3 min at 353 K. This difference reflects the key role acceptor 1,4-dicyanobenzene (DCB, Scheme 4) have been
of both the temperature and aluminum content in the DHP•+ introduced into zeolite-Y cavities and their photoevents have
electron recombination process.353 When trans-stilbene (t-St, been studied.360 Upon excitation of the composite, a caged
Scheme 4) is trapped in zeolite cages, it also has a long-lived An•+ radical-cation is formed, followed by an electron migration
CS, one that stays for hours.341,357,358 Within the channels of to the DCB molecules, which generates its radical anion
nonacidic-doped ferrierite, mordenite, and inverted mordenite (DCB•−). In zeolite-Y samples with different metal cations (Li+,
zeolites [M-FER, M-MOR and M-MFI, respectively where (M Na+, K+, Rb+, and Cs+), the electron migration is affected by the
= Na+, K+, Rb+ and Cs+)], a photoinduced intermolecular ET alkali metal in zeolite-Y framework.360 Transient diffuse
reaction was observed in t-St.341,358 In these composites, the reflectance experiments yielded three distributions of rate
back ET process includes direct charge recombination (CR), constants characterizing the generated pair of the radical’s
hole transfer (HT), and electron−hole recombination to cation and anion. The fastest behavior was observed for NaY
reform the caged t-St (Figure 26).341 sample, with three time constants (52 min and 9 and 75 h),
The CS state lifetime depends on the caging zeolite (FER, while the slowest one was recorded for RbY (220 min and 22
MFI, and MOR) and the nature of the extra framework cation. and 108 h).360 Thus, NaY is found to be the best zeolite to
When t-St•+ is trapped in the K-FER, K-MFI, and K-MOR facilitate charge separation by electron migration in photo-
zeolites, its lifetime decreases from 10 h to 20 min and 0.4 s, excited species. Another report studied ET processes between
respectively.341 This difference is related to a restriction to the electron-donor p-terphenyl (PTP, Scheme 4) and acceptor
move in these hosts and thus governed by the size of the cages, DCB within Na-ZSM5 zeolites.357 In the absence of DCB,
being smaller for FER (pore size dimensions = 0.42 × 0.52 photoexcitation of PTP in PTP/NaZSM-5 composites
nm2) than for MFI (0.53 × 0.56 nm2) and MOR (0.67 × 0.76 generates PTP•+ radical-cations with a lifetime of 21 min that
nm2). In FAU, zeolites having a large supercage (1.3 nm in evolve into an electron−hole pair in which the electron is
diameter), the t-St•+ lifetime is 150 μs.359 Interestingly, the CS captured by the zeolite framework with a time constant of 653
AA DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

min (Figure 27).357 Both species are more stabilized in the situation was observed for Ru(bpy)32+ (acting as D) and
presence of DCB molecules.357 In PTP-DCB/NaZSM-5 methylviologen (MV2+, Scheme 4) (acting as A) that were
encapsulated in clay particles of different sizes.361 ET was
observed only in the presence of small clay particles. A long-
lived CS state was also detected in the caged [(bpy)2RuLDQ]4+
complex (being LDQ = 1-[4-(4′-methyl)-2,2′-bipyridyl]-2-[4-
(4′-N,N′-tetramethylene-2,2′-bipyridinium)]ethene)) and ad-
sorbed MV2+ in zeolite.363 In this system, the long-lived state
results from a competition between the slow back ET reaction
(rate constant = 3.0 × 105 s−1) and an electron migration that
moves the charge away from the surface of the zeolite, caused
by self-exchange between the densely packed MV2+ groups
(rate constant = 2.0 × 105 s−1).363 Therefore, strong
Figure 27. Illustration of the competitive electron transfer events of confinement effects and an internal electrostatic field contribute
PTP to DCB dye and ZSM-5 zeolite framework. Reprinted from ref
to the deceleration of the CR process and stabilize the
357. Copyright 2016 American Chemical Society.
photoinduced transient species, resulting in exceptionally long
lifetimes of several seconds to hours and even several days.
complexes, the radical-cation and electron−hole pair lifetimes 3.2.2.2. Electron-Donor Molecules Encapsulated in Silica-
are longer (83 and 3300 min, respectively). This long-living
Based Acceptor Materials. The encapsulation of organic
state is mainly due to the strong confinement that is imposed
molecules that have an electron D group in their structure
by the ZSM-5 nanochannels and the intrinsic strength of its A
can enable an ET process to the silica-based metal-doped
sites.357 In addition to that, both the ZSM framework and the
framework. Such a reaction was observed when 7-HQ (Scheme
encapsulated DCB molecules compete for the electron
capture.357 3) was encapsulated in titania-doped SBMs.200 The interaction
The introduction of metal complexes, acting as electron of 7-HQ with MCM41, TiMCM41 (6% Ti doped), and purely
acceptor, into SBMs has also been extensively stud- Ti-mesoporous (TiMeso) materials was studied at the fs−ns
ied.344,348,350,361−363 In these hosts, long-lived CS states with regime.200 First, 7-HQ interacting with regular MCM41 results
lifetimes from nanoseconds to seconds were re- in E forms that have different H-bonds to the host framework
corded.350,361,363,364 For example, when phenosafranine and a K tautomer formation of S0. Upon electronic excitation at
(PHNS, Scheme 4) was adsorbed in the Y-zeolite surface in the caged E form, proton transfers in these trapped H-bonded
the presence of caged metal complexes of nickel(II), (Ni- structures to the MCM41 framework results in confined
(bpy)3)2+ induced intermolecular ET processes in the nano- anions.200 These dynamics were discussed in section 3.1. These
seconds regime (Figure 28). Upon excitation of the composites, anion and keto forms can give an electron to an electron-A
the adsorbed PHNS molecules transfer an electron to the Ni partner in the host. ET was thus examined using MCM41
complex that is also encapsulated in the Y-zeolite supercage.364 doped with titania.200 No ET process was observed using
Similar behavior was found when tris(2,2′-bipyridine)- regular MCM41 host, but emission quenching due to ET was
ruthenium(II) (Ru(bpy)32+) is trapped in Y-zeolite that is recorded in TiMCM41 (titanium-doped material with 6%
surrounded by bipyridinium ions.350 In this complex, the loading) and TiMeso (crystalline mesoporous all-TiO 2
presence of bipyridinium ions increases the lifetime of the structure with a uniform pore size distribution similar to that
photoexcited Ru(bpy)32+, promoting the extent of the forward of MCM41) (Figure 29).200
ET reaction. Moreover, the longer distance between the In the TiMCM41 complexes, the ET process occurs in 0.3−5
encapsulated Ru(bpy) 32+ complex and the external bipyr- ps, while in TiMeso, the emission decays are even shorter
idinium ions retards the back ET reaction.350 Clearly, the (Figure 29, panels B and C).200 The interaction with TiMeso
distance between D and A molecules plays an important role in results in two different EI reactions: one from the anionic forms
this kind of reactions and can even cause its inhibition. This (0.1−0.4 ps) and another from the Z ones (0.8−1.2 ps).200 The

Figure 28. Representation of PHNS adsorption on the surfaces of (A) NaY and (B) (Ni(bpy)3)2+ NaY. Reprinted with permission from ref 364.
Copyright 2014 Springer.

AB DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 29. Femtosecond-emission transients of 7-HQ interacting with (A) R-MCM41, (B) Ti2-MCM41, and (C) TiMeso materials in
dichloromethane suspensions, gated at the indicated wavelengths. Reprinted from ref 200. Copyright 2012 American Chemical Society.

difference between EI times in the anion and Z structures is of titanium atoms (6%), and TiO2-MCM41, which is doped on
explained by the differences in excited state energies compared the surface with large nanoparticles, some of the trapped 7-HQ
with that of the titania conduction band (Figure 30). The molecules are not in contact with the titania domains, resulting
in different population not all having ET with the titania sites.
However, for TiMeso materials and neat P25 NPs, which
directly contact the 7-HQ, the efficiency of the ET reaction is
higher.200
The photosensitization of TiO2 in cages or micro- and
mesoporous materials has been used as a strategy for the
formation of ET complexes. Several studies of the photo-
physical properties of PHNS (Scheme 4) adsorbed on different
zeolites and mesoporous materials that were photosensitized
with TiO2 NPs have been published.230,233,364−367 PHNS is an
azine derivative that can behave as an electron A and D in both
the ground and excited states, which makes it an important
molecule in the field of ET. When this dye is trapped within
NaY and ZSM-5 supercages, its fluorescence decays are
biexponential, with time constants of 0.48 and 1.42 ns
(PHNS/NaY) and 0.13 and 0.55 ns (PHNS/ZSM-5), which
are similar to those observed when silica nanoparticles were
used, reflecting that this molecule mostly interacts with the
zeolite surface.365 In both complexes, the presence of TiO2 NPs
Figure 30. Schematic representation of energy levels of S0 and S1 of causes a static emission quenching that is proportional to the
AN and Z forms of 7-HQ interacting with TiMeso material in a TiO2 concentration in the host.365,366 When TiO2 NPs are in
dichloromethane suspension. The scheme shows the values of time MCM41 material, they also enable an EI process from the S1
constants of EI to titania conduction band. Reprinted from ref 200. state of PHNS to the CB of the semiconductor.230 However,
Copyright 2012 American Chemical Society. the ET reaction is more efficient in the MCM41-based
complexes than in the zeolite-based ones since the
encapsulation of the dye inside of the larger MCM41 pore
excited anions emit at wavelengths shorter than the Z emit, so
allows greater contact with the TiO2 NPs, thus producing faster
faster emission quenching of the formers was observed. When
EI processes.230,366 Similar studies were reported for other dyes
7-HQ interacts with TiO2 nanoparticles (P25), it behaves like
in TiMeso, reflecting the strong interaction of this guest with such as PF, Th+ (Scheme 2), and methylene blue (MB)
the titania surface of the MCM41 materials and the small (Scheme 4) interacting with ZSM-5 and NaY zeolites and
confinement effect on the EI process.200 P25 nanoparticles MCM41 materials.230,232−234 The results show that the
were also mixed with MCM41 materials.200 In this case, the 7- photophysical properties of these dyes are affected by the
HQ interacting with the TiO2/MCM41 hybrid material showed morphology and the composition of the host materials and the
the same dynamics that it was recorded in MCM41.200 Due to presence or absence of TiO2 NPs.230,232−234 In SBMs, the
the large size of the used titania nanoparticles (50 Å), they are fluorescence of these molecules decay biexponentially; the short
located outside of the MCM41 nanochannels, where caged 7- component is attributed to the molecules that exhibit strong
HQ barely has contact with them.200 This spatial arrangement interactions with the silica host, and the long one is due to
indicates that the ET processes in these composites needs a molecules whose environment is similar to the solution.232,234
direct contact between caged 7-HQ and the titania domains. The introduction of TiO2 NPs inside these hosts produces a
For the cases of TiMCM41, which is doped with a low amount decrease in the fluorescence intensity attributed to the
AC DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 5. Molecular Structures of the Dyes Interacting with SBMs and Discussed in the Energy Transfer Section (Section 3.3)

photosensitization of TiO2 by caged dye molecules through the the host can be controlled to change the photobehavior of the
EI process.232,234 hybrid complexes.
The reports reviewed in this section reflect how the 3.3. Energy-Transfer Events
molecular encapsulation of dyes in SBMs improves their ET-
The EnT process, in which an excited D molecule transfers its
based photoreactions by increasing the stability of the
energy to A, can occur via FRET (dipole−dipole interaction)
photoinduced CS state and preventing the back ET reaction and/or via the Dexter mechanism (electron exchange). A large
or making it slower. The results suggest strategies for better amount of works on this topic has been performed with
designs of trapped electron D−A systems that might be used in chemical and biological systems.5,16,94,368−380 EnT occurs in
solar cells, photocatalysis, and molecular memory systems and both natural and artificial systems. Several review articles have
in nanophotonics in general. Some of these approaches were examined EnT in solution,32,287,380,381 in biological systems,
used in the prototype devices described in the application part and hybrid materials such as zeolites, clays, and polymeric
(section 5). These strategies indicate that proper selection of nanoparticles.2,4,5,7,16,382−384 Recently, the photochemical and
the SBMs pore size and semiconductor particle diameter can photophysical aspects of EnT in liquid solutions and hybrid
tune the interaction in the composites toward a desired goal materials have been reviewed.16,32 Ultrafast photochemical
such as photosensitization or stronger fluorescence of the reactions in homogeneous liquids including intra- and
confined system. In addition to this strategy, the concentration intermolecular charge-transfer processes, the design of optical
of the dye and, thus, the possible formation of aggregates within artificial antennas using hybrid materials, dye-doped latex
AD DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

nanoparticles, and molecular antennas based on D and A dyes molecules to externally absorbed PyB+ molecules is the
were discussed.16,32 As we have already commented on, dyes dominant pathway (Figure 31).399
(D and A) can be adsorbed on the silica surfaces or within their
channels, in which the cavities of microporous and mesoporous
structures provide a D−A distance that ranges from 20 to 100
Å. At these distances, FRET processes span from the
femtoseconds to the milliseconds regime.16,88,370,375,376 To
the best of our knowledge, EnT in all the systems involving
SBMs follows FRET to occur, without any information on the
possibility of the Dexter mechanism. The confining space
afforded by the nanochannels of the mesoporous materials and
most of the zeolite structures allows dyes to orient and pack
when being encapsulated, thus promoting ways to “control”
their relative orientation and distance between different D and
A units for FRET. This consequently allows the creation and
tuning of photonic antennas that are based on
EnT.5,16,374,376−378,385−389 The related process can also
manifest when the excited encapsulated partners directly Figure 31. Cartoon showing the difference in the energy transfer ways
interact with the silicate framework, thus affecting the efficiency in (A) (v-PyB+/L-g)AP from internally incorporated PyB+ molecules to
of other competitive processes. externally absorbed ones, and (B) (v-PyY+/L-g)AP mostly between
In other sections of this review (related to PT, ET and internally incorporated PyY+ molecules of different orientations in the
aggregates formation) that are included in section 3, the channels of zeolite L. Adapted with permission from ref 399.
Copyright 2014 Elsevier B.V.
systems under study have been classified according to the
properties of the molecules; in contrast, the EnT part has been
divided according to the materials that were used. The reason On the other hand, the EnT in vertically aligned PyY+-
for this classification lies in the complexity of the systems containing zeolite L monolayers (v-PyY+/L-g)AP occurs mostly
described here; separating classes of molecules would between internally incorporated PyY+ molecules that have
complicate the comment and review of the published works. different orientations in the zeolite L channels. This transfer
3.3.1. Energy Transfer in Dye-Doped Zeolites. results in an increase in the fluorescence depolarization ratio
Encapsulation of organic dye molecules inside the channels of (I⊥/I∥).399 Thus, the EnT process can be controlled by
zeolite materials enhances the stability of the dye molecules selecting appropriate molecules. Another class of hybrid
and, in many cases, prevents aggregation and fluorescence assemblies that were based on zeolite L has also been
quenching.72,94,385,387,390−393 One of the most used zeolites in explored.375 In these systems, the A molecules (Ox+, Scheme
the field of EnT is zeolite L. This material possesses a 2) are organized in the channels of zeolite L, while the D
crystalline structure that is formed by one-dimensional species, CdSe/ZnS QDs, are adsorbed on the surface. The QDs
hexagonal channels of 7−13 Å in diameter, making it an absorb light and undergo electronic EnT, funneling the energy
appealing host for organizing highly fluorescent dyes.394,395 inside the channels and providing an artificial antenna. The
Geometrical constraints imposed by its pores lead to a mechanism for this photoinduced process is ascribed to
supramolecular organization of the guests that is ideal for vectorial FRET and was demonstrated and quantified using
fundamental studies in unidirectional EnT.5 A large number of microscopy techniques, steady-state, and nanosecond time-
studies based on the organization of dye molecules trapped in resolved spectroscopy.375 The EnT efficiency was determined
its channels has been published.94,374−376,385,387,394−396 Some of to be 70%, with a Förster distance (R0) of 73 Å and rate
these studies have reported on the introduction of “stopper” constant (kEnT) of 1.65 × 108 s−1. These findings show that the
molecules that control the interaction between the zeolite- system is efficient in absorbing light and vectorially transferring
encapsulated dyes and the outside environment.370,389 These the electronic excitation energy to an acceptor located inside
stoppers at the gates of the pores are selected to act as blockers, 1D nanochannels.375
not allowing the trapped molecules to exit. They can also be Another strategy to get EnT process in zeolite materials was
photoactive systems such as energy donors, which, in demonstrated in an inorganic chalcogenide-based semiconduc-
combination with other molecules trapped inside the channels tor zeolite material (denoted RWY) that served as an UV−
that act as energy acceptors, creates an efficient antenna visible light-harvesting host.386,397 This optically active
system.370,386,388,389,395,397,398 To study the EnT in zeolite chalcogenide-based zeolite, capable of integrating porosity
crystals, microporous zeolite L and two dye molecules, with semiconductivity and photoluminescence, provides a
pyronine B and pyronine Y (PyB+ and PyY+, respectively, new opportunity for framework-involved EnT between an
Scheme 5), were used as the host and guests, respectively.399 inorganic host and an organic guest. The initial multistep
Polarized-fluorescence microscopy images have demonstrated vectorial EnT scaffold was fabricated by encapsulating acridine
that the incorporated PyB+ molecules were forced to align orange into the RWY porous framework and then covering the
along the direction of the channel of the zeolite L crystal and formed capsules with the same dye and rhodamine B molecules
that the arrangements of the PyY+ ones were located mostly at (AO and RhB, respectively, Scheme 5).397 Upon excitation of
both ends of the channel. The study analyzed the dominant the RWY host, it transfers its energy to AO molecules and then
EnT between molecules that were internally incorporated into to RhB ones to give rise to a visible-light emission. The
and externally adsorbed on the zeolite L monolayer.399 For the photodynamic behavior was monitored getting the emission
case of vertically aligned PyB+-containing zeolite L monolayers decays of the RWY host, RWY/AO, and Rhb-(RWY/AO)
(v-PyB+/L-g)AP, the EnT from the internally incorporated PyB+ composites. Multiexponential behavior with three time
AE DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

constants was observed for the three samples.397 The average that are not observed in solutions. Some D and A molecules in
fluorescence lifetime (3.50 ns) in RWY is longer than that in solutions do not exhibit EnT, while, when trapped into the
the presence of AO molecules (3.08 ns for RWY/AO pores, they show the intermolecular energy events. For
complexes), indicating EnT from RWY to AO guest. Similar example, TPC1 (Scheme 4) fluorescence behaves normally in
behavior was observed when RhB molecules were added to the solution, but upon encapsulation in MCM41 pores, it behaves
above composite. Thus, two EnT events, from RWY to AO differently.323 We commented on this study in the ET section,
(EnT1) and from AO to RhB (EnT2), occurred with relatively as TPC1 molecules transfer electrons when interacting with
low efficiencies (12% and 9%, respectively).397 A multistep (or titanium inside host materials (TiMCM41 and TiSBA15). In
stepwise) EnT assembly based on the RWY framework was dichloromethane solutions, two fluorescence lifetimes (800 and
further developed by encapsulating both pyronine and 2.5 ps) were observed, assigned to neutral and anion forms,
proflavine ions (Py+ and PFH+, Schemes 2 and 3, respectively) respectively. When TPC1 is trapped in MCM41, an additional
(Figure 32).386 The UV photons that were absorbed by the small and shorter component (180 ps) was observed and
attributed to a self-quenching by EnT.323 The origin of this
process was proposed to be singlet−singlet annihilation, which
originates from EnT between two excited TPC1 molecules.323
The strength of the interaction of a trapped guest with the
OH groups of the framework surface and the nanosized pores
of its channels can also alter the EnT between encapsulated
dyes. This has been demonstrated in a study in which the
FRET efficiency between Rhodamine 6G (R6G, Scheme 5),
acting as donor, and RhB as acceptor, was enhanced due to the
variation in the strength of interaction of both partners with the
MCM41, MCM48, and SBA15 mesoporous materials.400 The
FRET efficiency between the dyes was determined using the
ratio of the emission intensities (I0/I) in the absence and
presence of quencher (A) at different concentrations. From the
slope of the plot of I0/I (emission quenching) versus RhB
concentration, Stern−Volmer quenching constants (Ksv) were
obtained. The analysis yielded Ksv = 6.11 × 105 M−1 for
MCM48, 2.92 × 105 M−1 for MCM41, and 2.31 × 105 M−1 for
SBA15.400 Therefore, the highest EnT efficiency was observed
Figure 32. Multistep energy-transfer processes in RWY/PFH+/Py+
for MCM48 composites, followed by MCM41 and SBA15
host−guest antenna systems. Reprinted with permission from ref 386. ones.400 The EnT is enhanced in these mesoporous materials as
Copyright 2015 Royal Society of Chemistry. indicated by the larger Ksv values when compared to those in
water with a Ksv = 1.65 × 105 M−1. The three-dimensional
structure of MCM48 is the most favorable for the FRET
RWY host were transferred to the PFH+ ions and then to Py+ process since its channels can highly disperse the dye
ones to produce a visible-light emission. The treatment of the molecules. On the other hand, the encapsulation of bulkier
guest ions after their encapsulation such as acidification of the dye molecules can result in pores obstruction in MCM41, thus
PFH+ ions and the solvation of the guests were also decreasing the EnT efficiency. Although SBA15 has a large pore
investigated to tune the EnT efficiency in such host−guest diameter (d ∼ 60 Å), which may be advantageous for the
antenna systems. The acidification of the PFH+ guest dispersion of the guest molecules, the EnT efficiency was lower
dramatically enhances the EnT efficiency in both steps, than in MCM41 (d ∼ 25 Å).400 The authors explained this
increasing from 14 to 33.7% for RWY-PFH+ (EnT1) and difference as a result of a lower density of OH groups in SBA15,
from 29.1 to 37.8% for PFH+-Py+ (EnT2).386 On the other which leads to a longer distance between the guest
hand, solvation experiments showed that the EnT efficiency molecules.400 The EnT efficiency can also be controlled by
depends on the used solvent. For example, while water is not modifying the framework properties with the introduction of
beneficial to EnT, organic solvents such as ethanol, metal atoms (doping), as we have commented on for other
tetrahydrofuran, and toluene drastically promote the EnT processes. The effect of doping was demonstrated for MCM48
from RWY to PFH+ (EnT1) and from PFH+ to Py+ (EnT2).386 materials having Al and Ce atoms in their frameworks.400 The
These results indicate that in addition to the structure of the presence of these metal atoms decreases the FRET efficiency
guest, the environment plays an important role in controlling because of an increase in the ability of the material to accept an
the EnT in these host−guest antenna systems. electron from the D molecule (R6G).400 EnT in Ce-MCM48 is
3.3.2. Energy Transfer in Dye-Doped Mesoporous lower than in AlMCM48, and it is less efficient in both than in
Materials. When organic dyes are encapsulated within regular MCM48. The results also demonstrate that EnT
mesoporous materials, efficient EnT systems can also be between caged dyes can be regulated by doping the framework
formed. The ability of these hosts to encapsulate a variety of with metal atoms of different oxidation abilities (here Al and Ce
aromatic molecules along with the ease of chemically modifying atoms).400 Adjusting these properties enables different colors of
the silica framework make them the hosts of choice for the emission from these composites to be observed, which could be
fabrication of nanoantenna-based composites. The encapsula- of potential interest in photonics and sensing.
tion can significantly stabilize the D and the A entities, thus 3.3.2.1. Systems with Donor Dyes Bound to a Mesoporous
giving rise to more efficient FRET. The smaller space in the Surface. Another strategy used to construct host−guest
nanochannels of the materials can promote molecular events antenna systems based on SBM composites is to directly
AF DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 33. (A) Chemical structures of TPPy-derivative organosilane precursors TPPy-Si and PBI dyes 1−4. (B) Emission quenching efficiency of the
P123-TPPy films doped with PBI dyes. Reprinted from ref 378. Copyright 2012 American Chemical Society.

dope the donor unit into the mesoporous material framework spectra of the no doped Bp-PMO composites showed the
or covalently link it to the inner walls, while the acceptor is formation of three types of excimers (E1, E2, and E3), each with
encapsulated within the nanochannels. Dye-doped periodic different formation time constants.401 Upon excitation of Bp-
mesostructured organosilica films with different host−guest PMO, its fluorescence spectrum up to 100 ps has a peak at 330
arrangements were fabricated and examined.378 The photo- nm, assigned to E1 (Figure 34A). From 0.1−0.2 ns to
functional system is formed by fluorescent tetraphenylpyrene approximately 20 ns, the fluorescence spectrum becomes
(TPPy)-silica mesostructured films, which act as the host broader, and the change is attributed to the formation and
donor, and perylene bisimide (PBI, Scheme 5) dye, which acts emission of E2 and E3 (Figure 34A). Interestingly, in the
as the guest acceptor (Figure 33A). In these composites, the presence of 0.1 mol % C1, the rapid formation of Bp excimers is
effect of different substituents in the PBI molecular structure on followed by an EnT to C1 in the mesochannels (Figure 34B).
the EnT process has been explored (1−4 in Figure 33A).378 From 0.1 to 0.2 ns gating time of its fluorescence spectra, a new
In dye 1, weak fluorescence quenching due to aggregates band at 440 nm was recorded. The emission intensity of this
formation inside the mesostructured material was observed band increases with time, while that of E1 decreases, indicating
(Figure 33B). Here, the EnT process occurs between the TPPy an EnT from the excimers of Bp to C1. In the case of larger C1
moieties and the PBI dye aggregates. For dye 2, which is a PBI doping (0.36 mol %), spectral changes due to excimers
derivative with bulky substituents and polar anchoring groups, a formation were also detected until 0.2 ns (Figure 34C). After
low population of aggregates was formed, and the monomers on, the decay of the 365 nm emission band and rise of the 440
were the predominant species. The composite exhibits an nm one were shorter than those in 0.10 mol % C1-doped Bp-
efficient FRET from the host TPPy to the PBI dye. Using dye PMO (Figure 34, panels B and C). The excimers fluorescence
3, a mixture of monomers and aggregates species was recorded, (390 nm) increased with a decrease in the monomer
although self-quenching by face-to-face association was sup- fluorescence (300 nm), after which the C1 fluorescence (465
pressed due to the bulky R1 and R2 substituents (Figure 33A). nm) increased with a decrease in the excimer fluorescence.
However, the FRET process was not observed since the D−A These results clearly indicate that the S1 state of the Bp moiety
(host−guest) distance was longer than the critical Förster prefers to form excimers rather than to transfer its excitation
radius (ca. 4.5 nm). This host−guest arrangement is useful in energy to C1 (although these two processes are competitive)
certain cases (e.g., photoluminescent materials) because the and that EnT occurs after excimer formation. The Förster
undesired EnT is completely suppressed due to the isolation of radius between both sets of partners was calculated to be 3.2
D and A partners. Remarkably, dye 4 exhibits the most efficient nm, suggesting that other mechanisms contribute to the
fluorescence quenching, which is caused by a photoinduced ET observed efficient EnT. Monte Carlo simulation suggested
from the host, TPPy-silica, to the guest, PBI (Figure 33A). The that FRET occurs from three types of Bp excimers to C1
transient absorption experiments showed that radical (PBI•−) molecules that are located in the vicinity of the pore surface. It
generation occurs in <100 ps, and the resulting CS state lived happens at the hydrophilic silica layers on the pore walls, while
for nanoseconds (Figure 33B). Therefore, dyes 2 and 4 suffer some of the Bp excimers are not affected by any A.401 These
from efficient emission quenching but as a result of different results demonstrate how the heterogeneity of the system affects
competing processes (EnT and ET, respectively).378 These the EnT mechanism, with an emphasis on how competitive
results further demonstrate the relationship between the EnT ultrafast reactions such as excimer or aggregates formation
efficiency and the guest molecular structure. directly contribute to the efficient EnT process.
Another system with comparable architecture that exhibits Mesoporous organosilica materials have also been used as
EnT comprises biphenylylene-bridged periodic mesoporous hosts for EnT systems.402 FRET was reported between 1,8-
organosilica (PMO) powder doped with coumarin 1 (C1, naphthalimide covalently attached to the host and trapped
Scheme 5).371,401 The picoseconds time-resolved fluorescence N,N′-bis(2,6-dimethylphenyl)-3,4,9,10-perylenediimide (NI
AG DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

was observed and attributed to aggregate species.402 The


Fö rster radius (R0) was found to depend on sample
concentration, being, respectively, 2.5 and 3.1nm in diluted
and concentrated ones, and the EnT process is less efficient in
the former. Nanosecond time-resolved fluorescence experi-
ments were performed in concentrated NI (D) samples at
different concentrations of PDI (A). In its absence, the sample
exhibits biexponential behavior with time constants of τ1 = 6.6
ns and τ2 = 35.9 ns, assigned to the emission of monomer and
aggregate NI species, respectively.402 However, in the presence
of 9% mol of PDI, the sample displays a triexponential behavior
with time constants, τ1 = 2.2 ns, τ2 = 9.6 ns, and τ3 = 48 ns. The
shortest component reflects the EnT process from NI to the
PDI, while the other time constants are due to the emission of
monomer and aggregate NI species, respectively.402 For these
composites, the emission spectra span the whole visible
spectrum, making them strong candidates for white-emitting
LEDs.
3.3.2.2. Systems with Acceptor Dyes Bound to a
Mesoporous Surface. Another approach to making artificial
D−A antennas is binding the acceptor entity to the outside
surface of the silica-based material. A hybrid photostable energy
D−A mesoporous SBA15 silica system was made in this
manner (Figure 36).377 In this composite, the energy D, Super

Figure 34. Time-resolved fluorescence spectra of (A) no doped, (B)


0.10 mol % coumarin-doped, and (C) 0.36 mol % coumarin-doped
BpPMO. Inset: fluorescence rise and decay curves of the sample in
(C). Reprinted from ref 401. Copyright 2013 American Chemical
Society.

and PDI in Scheme 5 and Figure 35).402 The optical properties


of these complexes could be tuned depending on the
concentration of NI in the composite. At lower concentrations
of NI (D), a blue emission was recorded and assigned to
monomers, while using a higher NI loading, a green emission

Figure 36. (A) Schematic representation of the synthetic route for the
studied hybrid antenna. Digital photographs under UV light of (a)
SY@SBA-15, (b) IRIS3@SBA-15, and (c) IRIS3-SY@SBA-15 in (B)
2-propanol suspensions and (C) powder. Reprinted with permission
from ref 377. Copyright 2012 Wiley-VCH.

Yellow (SY) polymer, is encapsulated in SBA15 pores, while


the A, IRIS3 cyanine dye, is covalently linked to the outside of
the framework (Scheme 5). Time-resolved fluorescence studies
showed that the SY/SBA15 complexes behave differently in the
absence and presence of IRIS3. While that of SY/SBA15 decays
biexponentially (τ1 = 1.1 ns and τ2 = 6.8 ns) with an average
lifetime of 3.5 ns, the fluorescence of IRIS3-(SY/SBA15) is
monoexponential (2.1 ns), indicating that FRET occurs from
Figure 35. Cartoon showing PDI-doped SBANI materials and FRET the caged SY polymer to the externally linked IRIS3 cyanine
from NI to PDI dyes. Reprinted from ref 402. Copyright 2015 dye.377 The FRET efficiency, R0, and kEnT were calculated using
American Chemical Society. the emission lifetimes to be 40%, 37 Å, and 1.90 × 108 s−1,
AH DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 37. (A) Structure composition of coumarin/DOC/phyllosilicate hybrids. (B) Cartoon of 7-hydroxy-4-methylcoumarin and DOC molecules
confined to the interlayer of a phyllosilicate. The arrows show the energy transfer from coumarin to adjacent DOC via ultrafast FRET and from
remote coumarin to DOC via radiative transfer. Reprinted from ref 408. Copyright 2010 American Chemical Society.

respectively.377 The high efficiency of EnT in this hybrid system without Au NPs but 1.7 ns in the Au-zeolite composites. The
is due to the large spectral overlap between the D emission and presence of Au NPs also was accompanied by a decrease in the
A absorption. On the other hand, the estimated D−A distance fluorescence quantum yield. Estimated efficiency of EnT from
is consistent with the thickness of the silica wall of SBA15 (33 the dye to the Au NPs is 70%. Using the SET method, the
Å), confirming that the IRIS3 dye is anchored on the external distance between C480 and Au NPs attached to zeolites was
surface of the material and not inside its pores.377 estimated to be approximately 70 Å.
Similar nanoscale architectures in which R6G was encapsu- 3.3.3. Donor/Acceptor Concentration Effects on
lated in the channels of MCM41 with different amounts of gold Energy Transfer. The role of the relative concentrations of
nanoparticles (Au NPs) anchored on the silica surface have energy D and A partners is also an important factor to consider
been designed.403 The photoluminescence intensity of the in EnT systems. Homo EnT (A and D have the same molecular
caged R6G dye is quenched by Au NPs. The quenching structure) is described in the section on aggregates. Here, we
efficiency (76, 64, and 27%) depends on the concentration of will focus on hetero transfer (A and D have different molecular
the Au NPs [1:0.05, 1:0.1, and 1:0.2 Au:(R6G/MCM41) ratios, structures). In a system containing fluorescein (D) and RhB
respectively].403 The fluorescence lifetimes behave similarly, (A) in silica materials (Scheme 5), two types of dye-silica
resulting in average decay times of 1.85, 2.16, and 3.17 ns for nanoconjugates have been designed and examined: one based
Au:(R6G/MCM41) ratios of 1:0.05, 1:0.1, and 1:0.2, on silica nanoshells (SNS-dye) and another on silica nano-
respectively, while the decay lifetime for the R6G/MCM41 particles (SNP-dye).406,407 In both nanosystems, the concen-
composites without AuNPs is 3.90 ns.403 Therefore, the tration of donor molecules affects the FRET efficiency and
decrease in the R6G photoluminescence intensity and the dynamics. While the amount of acceptors was kept constant,
decay times in the presence of Au NPs confirm the occurrence the D:A concentrations ratio changed from 0.5 to 3 in SNS-dye
of EnT. Its efficiency from the trapped R6G molecules to the and from 0.5 to 4 in SNP-dye systems.406 Femtosecond
external Au NPs are 52, 44, and 17% for the R6G/MCM41 fluorescence experiments showed that when the SNS-dye is
composites that contain 0.05, 0.1, and 0.2 relative amounts of excited at 520 nm, a residual D emission was observed at all
R6G, respectively.403 As the number of Au nanoparticles values of the [D]:[A] ratio. However, when the SNP-dye was
relative to that of the R6G/MCM41 composites increases, the excited, the D emission was quenched systematically from [D]:
narrowing of the EnT channel becomes more evident. This [A] = 0.5:1 to 2:1, leaving only residual emission in the case
decrease in the EnT efficiency is due to the formation of Au when [D]:[A] = 4:1. These results indicate that EnT in SNP-
NPs aggregates induced by a self-assembly on the MCM41. dye is more efficient than in SNS-dye.406 When SNS- and SNP-
This aggregation produces a new plasmonic band in the long- dyes are excited at 630 nm (A absorption region), a long
wavelength region of the absorption spectrum, consequently lifetime component from the acceptor was observed, reflecting
reducing the required spectral overlap.403 The distances (r) a less efficient EnT process than that at 520 nm. Moreover,
between the D and the A (r = 79, 85, and 119 Å for 1:0.05, when the mixtures were excited at the A region, differences
1:0.1, and 1:0.2 Au:(R6G/MCM41) ratios, respectively) were between the SNS-dye and SNP-dye emission transients were
estimated by the surface energy transfer (SET) method.403 This recorded. For the case of SNS-dye, a risetime of 500 fs in the
way is useful in calculating long EnT distances, which is not emission signal of [D]:[A] = 0.5:1 to 2:1 confirmed the
possible with the FRET process because its upper limit distance occurrence of FRET, while for [D]:[A] = 3:1 and all [D]:[A]
is 80 Å.403 Finally, anisotropy decay experiments revealed that ratios with SNP-dye, no risetime was observed.406 The absence
R6G molecules are aligned into the channels of the mesoporous of the rise time signal is due to an ultrafast and efficient FRET
silica, and that the EnT process is unidirectional.403 Therefore, in the SNP-dye system that is not happening in the SNS-dye
supramolecular organization inside and outside the channels one. For example, the FRET efficiency in SNS-dye and SNP-
leads to more efficient light harvesting systems. dye with a [D]:[A] = 2:1 ratio, was 91 and 97%, respectively.
Similar studies were carried out using C480 (Scheme 4), The difference was explained as a result of the different
confined in Y-zeolite functionalized with Au NPs.404 The methods of synthesis of both nanoconjugates.406 For the case of
fluorescence anisotropy decay of C480 increased from 100 ps in SNP-dye, the synthesis leads to dye-rich core formation which
bulk water to 770 ps in zeolite Y due to the restricted rotation results in dye molecules with a high proximity, generating
of the dye imposed by the confinement of the zeolite cavity. FRET in a more efficient way than in the SNS systems. On the
C480 can undergo easy twisting motion, leading to faster other hand, an increase in the D concentration decreases the
emission depolarization.306,311 Such a phenomenon has been FRET efficiency of both complexes (60 and 66% for [D]:[A] =
explored in coumarin 30 (C30, Scheme 5), using chemical 3:1 in SNS and [D]:[A] = 4:1 in SNP systems) due to
(CDs) and biological (HSA protein) cavities.405 The average competitive self-quenching of D molecules.406,407 Thus, the
fluorescence lifetime of coumarin 480 was 5.7 ns in zeolite union of the D and A molecules with these nanoconjugate
AI DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

systems forms FRET nanostructures that are highly efficient, t1/3 (t is time) are superimposed on the D’s radiative decay. The
but an optimum concentration of both partners is desired to contributions of these components in the emission signal
achieve a maximum EnT yield. become larger with increasing A content. However, for the 3D
Another system where efficient FRET depends on the model, the component decays as t1/2, and this contribution is
molecular concentration is formed by coumarin (7-hydroxy-4- smaller than that observed for the 2D one.408 For this
methylcoumarin, Scheme 5) acting as D and cyanine (3,3′- composite, the association of A and D partners originates from
diethyloxacarbocyanine, DOC, Scheme 5) acting as A confined an intercalation process in the synthesis that incorporates the A
in the interlayers of aluminum silicate nanofilms (Figure molecules by diffusion. During this intercalation, the molecular
37).408,409 In this case, the A concentration is changed relatively flow was limited to 2D motion inside the layer. As a
to that of D by using different [DOC]/[host hybrid] ratios (x = consequence, the 2D model, used to determinate the FRET
0.1, 1.0, and 10 mmol/100 g). The fluorescence spectra change parameters, is more accurate than the 3D one.408 The obtained
with the DOC concentration. At x = 0.1, the D and A FRET efficiency depends on the A concentration, which is 40,
molecules have similar emission intensities, while at x = 10, the 70, and 90% for x = 0.1, 1.0, and 10, respectively. An increase in
D emission decreases.408 Time-resolved photoluminescence A concentration favors their proximity with D, which are
experiments showed the existence of two different EnT confined in the interlayer of the silica, enhancing the efficiency
processes: an ultrafast FRET (in less than 5 ps) from coumarin of the 2D energy-transfer process (Figure 38). It has been
to an adjacent DOC, and a slower process (2.2 ns), which is a proposed that a larger amount of A and D can enable other
radiative transfer from distant D and A entities (Figure competitive processes such as self-quenching, dye aggregation,
37).408,409 Homo-energy transfer (Homo-EnT), and collective excitation
These results reflect the inhomogeneity in the distribution of effects due to exciton formation.408
the dye molecules in the composite. The fast FRET originates 3.3.4. Energy Transfer in Silica-Coated Quantum Dots
from short-range interactions between D and A molecules, and Metal Nanoparticles. Silica overlayers have been
which are more common at higher A concentrations, while the frequently used as protective or insulating materials deposited
slow process is associated with the emission and reabsorption on various nanostructures, especially semiconducting quantum
of photons (radiative transfer).408 To calculate the FRET dots (QDs) or metal nanoparticles. QDs possess remarkable
parameters, two- and three-dimensional (2D, 3D) cases were absorption and emission bands that are highly tunable upon
considered. For the 2D case, R0 is considered to be longer than shifting their diameters below the Bohr exciton radius, typically
the layer-to-layer thickness, while for the 3D one, the opposite in the single nanometer range.21,411 At the same time, various
situation is assumed because R0 of the D−A pair is in the range additional recombination processes occur that are not observed
of 2−9 nm.408,410 Figure 38 shows a comparison between the in the bulk crystals, with possible modifications helped by the
2D and 3D models in addition to the experimental decays. For presence of the silica coating. Metal (especially gold and silver)
the 2D scenario, rapid components that decay proportionally to nanostructures exhibit surface plasmonic resonance, which can
greatly enhance both the absorption and emission properties of
neighboring chromophores, with potential applications in, for
example, solar cells, LEDs and bioimaging.21,412,413 However,
overly close contact between the chromophore and metal
particle might result in unwanted efficiency quenching, and
therefore, the silica layer can act as a spatial separator.
A few such studies also investigated the photoinduced
dynamics occurring on the ultrafast time scale. CdSe nano-
crystals were coated with silica shells and coupled to Au NPs.414
The average fluorescence lifetime of the silica-coated nano-
crystals (emission spectrum peaking at 600 nm) not interacting
with Au particles was 18.4 ns. This lifetime decreased to 12.5
and 2.7 ns upon interaction with self-assembled films of Au on
different substrates. In the case of the shortest decay (2.7 ns),
the Au plasmonic band was more intense and better suited to
the emission band of CdSe QDs. At the same time, the CdSe
fluorescence intensity increased by a factor of 3 upon coupling
to Au NPs. This result was explained by a significant
enhancement of the excitation rate due to plasmonic resonance
induced by Au particles. Moreover, a strong blinking
suppression was also observed for the CdSe-Au system, with
an increase in the on-state fraction from 60% to more than
90%.414
The fluorescence enhancement of CdTe QDs interacting
with Au NPs was studied with different silica coating layer
thicknesses.415 The optimum SiO2 thickness was 4 nm, which
Figure 38. (A and B) Comparison of the observed donor (7-hydroxy- resulted in approximately 10 times more intense composites
4-methylcoumarin) emission decays at different acceptor concen- fluorescence than that of bare CdTe QDs (Figure 39). The
trations (c) and simulations for the 2D (C and D) and the 3D (E and photoluminescence time was decreased from 13 ns for the bare
F) cases. Reprinted from ref 408. Copyright 2010 American Chemical material to 2.8 ns for the optimum gold-silica composite, which
Society. confirmed an increase in the radiative decay rate of photo-
AJ DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 39. (A) Photoluminescence spectra and (B) emission intensity enhancement factor F for Au-SiO2 core−shell composites (ASC) of different
SiO2 thickness. (C) Emission decays of the bare CdTe and interacting with ASC of 4 nm. The inset shows the value of decay time constants.
Reprinted with permission from ref 415. Copyright 2013 The Optical Society.

Figure 40. (A) Illustration of the synthesis procedure of SiO2-coated CdTe quantum dots. (B) Biexciton Auger recombination lifetime vs
photoluminescence peak wavelength for CdTe cores (red ▲) and SiO2-coated CdTe quantum dots (blue squares); different peak positions
correspond to different size of CdS clusters. Reprinted with permission from ref 419. Copyright 2011 Royal Society of Chemistry.

excited fluorophores, attributed to the surface plasmonic silica NPs, the intensity of the dye photobleaching band was
resonance effect, due to an increase in the recombination rate greatly increased, but it suffered from a strong quenching with a
constant of electrons and holes. The enhancement factor 5 ps time constant. It was concluded that the plasmon-
decreased when the SiO2 thickness was above 4 nm. When the enhanced photocurrent generation could be observed only if
CdTe QDs were too close to Au NPs (silica shell of 1 nm), the the excited dye undergo ET into the titania in the first tens of
radiative decay rate decreases, while the nonradiative decay rate femtoseconds. After this time window, resonant plasmons were
was increased, resulting in emission quenching. Similar very likely to lose their coherence.418
observations were reported for Au nanorods coated with a The photoluminescence intensity of QDs is often decreased
silica layer and CdTe QDs attached to the silica surface.416 The due to the presence of Auger recombination. It is the primary
luminescence lifetime of CdTe decreases from 37 to 23 ns due exciton loss mechanism in colloidal nanocrystals not observed
to the increased radiative rate in CdTe upon interaction with in bulk crystals. A way to reduce this effect was proposed for
the Au nanorods, while the isolation of the silica layer prevents CdTe QDs coated with a silica layer.419 This shell contained
their quenching effect. small 1 nm CdS-like clusters, which were formed due to the
Composites of silica-coated gold nanorods and chlorine e6 dispersion of Cd2+ and thioglycolic acid in the silica shell
(Ce6, Scheme 5) photosensitizers were studied, motivated by (Figure 40A). The biexciton Auger recombination lifetime was
the search for new drug release methods.417 The mesoporous 5.4 ps for the bare CdTe QDs, while in the hybrid silica shell
silica shell contained amidogen so that the Ce6 molecules were structure, it increased to 19.2 ps. Moreover, the recombination
easily conjugated. The fluorescence lifetime of Ce6 was found lifetime increased with the size of the CdS-like clusters (Figure
to change drastically from 0.9 ns when attached to the gold- 40B). Nonradiative Auger recombination was also investigated
silica complex to 6 ns after its release. The release could be for giant CdSe/CdS QDs with additional silica coating.420 The
controlled by the power of the near-infrared (780 nm) observed splitting of exciton states was explained by the
continuous laser irradiation. Due to the strong absorption of modified charging environment due to the presence of the silica
Au nanorods at the surface plasmonic resonance band, the 780 shell. At low excitation fluences, the measured lifetimes were
nm laser beam could heat the system to release Ce6. The silica longer than 20 ns, and the charged exciton states showed
coating results in an enhanced and red-shifted surface strongly suppressed Auger recombination rates, resulting in
plasmonic resonance band. photoluminescence quantum yields close to 50%. However, at
Au-SiO2 core−shell NPs were studied in a dye-sensitized higher excitation fluences, the consecutive emergence of lower
solar cell system by an ultrafast transient absorption efficiency states was observed, indicative of higher-order
technique.418 A plasmon-enhanced light absorption of Z907 excitons.
ruthenium dye was observed, accompanied by an increased Much longer emission lifetimes (microsecond regime) were
photocurrent in the cells. However, the studies did not give a recorded for dense silicon nanocrystal ensembles fabricated by
direct conclusion on whether there was discrete EnT from the ion implantation into a SiO2 layer.421 The photoluminescence
plasmon mode to the excited dye state or if there was a direct quantum yield was measured as a function of silica thickness,
near field coupling. For the sample with the core−shell metal− reaching values as high as 60% at low excitation power. At high
AK DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 6. Molecular Structures of the Compounds Interacting with SBMs and Discussed in the Aggregates Formation Section
(Section 3.4)

fluences, the quantum efficiency was reduced, and internanoc- allowed.433−435 In contrast, J-aggregates exhibit absorption
rystal nonradiative decay processes were proposed as a possible bands at lower energies (S0→ S1 transitions).434,436 H- and J-
mechanism responsible for this decrease, in addition to aggregates have different dynamics than their monomer species.
biexciton formation followed by Auger recombination. Generally, for the aggregates, the emission lifetimes are shorter
3.4. Homo-Energy Transfer and Aggregates Formation than that of the monomers because of efficient deactivation
processes (e.g., EnT) that are caused by the molecular packing
EnT between the same dye molecules (Homo-EnT) is a special
of the monomers.433−436 Furthermore, for H-aggregates, this
case.369,373,422,423 It is strongly related to the formation of
emission quenching is high as the probability of radiative
molecular aggregates, which are favored in the confining space
transition from its S2 state is very low.434,437 The theory and
provided by the nanochannels of mesoporous materials and
zeolites. Thus, these two related phenomena (aggregation and experimental supports are well-documented and re-
homo-EnT) will be reviewed simultaneously. viewed.422,433−435,437−441 When the monomers are being
Many fluorescent dyes with planar molecular structures encapsulated within the nanomaterial pores/cavities, aggregates
consisting of three or more fused aromatic rings, such as formation can be either favored or inhibited, depending on the
rhodamines,424−426 pyronines,72,390,394,427 thionines,428,429 and molecular structure of the guest and the type of silica-based
oxazines,430−432 tend to aggregate at relatively low concen- material used.
trations. This molecular packing can result in formation of H- Hemicyanine dyes (Scheme 6) that were encapsulated in
(face-to-face) and J-aggregates (face-to-edge), which have nanosized pores of zeolites show this kind of behavior.442
different spectroscopic and dynamical behaviors. Classical Steady-state experiments indicated high fluorescence quenching
exciton theory predicts a splitting of the monomer excited as the concentration of the dye increased and the
state level into higher and lower exciton bands of the intermolecular distance shortened (from 12.6 to 2.1 nm).442
aggregates. But only one transition is allowed to give an The emission quenching was explained in terms of an
absorption band, which depends on the type of aggregates. H- intermolecular dipole−dipole interaction (FRET) between
aggregates are characterized by absorption bands at high the dye molecules (i.e., one excited hemicyanine molecule
energies (S0→ S2 transitions), and the S0→ S1 transition is not transfers its electronic energy to a hemicyanine in the S0 state).
AL DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 41. (A) View along the straight channels of the MAPO structures and (B) fluorescence images of PY dye in different MgAPO structures:
MgAPO-5, PY/AFI-H (left); MgAPO-36, sample PY/ATS-H (center); and MgAPO-11, PY/AEL-H (right). Reprinted with permission from ref 390.
Copyright 2013 Wiley-VCH.

Analysis of the time-resolved data showed that efficient EnT materials with a variety of structure types such as AFI, ATS, and
occurs when the intermolecular distance is longer than 5 nm AEL.390,427 The differences among these 1D structures mainly
(R0 = 2.2 nm). The emission decays were fit with a reside in their pore dimensions, which are 7.3 × 7.3 Å2 for AFI-
biexponential function. The shorter average time constant (MgAPO-5), 6.7 × 7.5 Å2 for ATS- (MgAPO-36), and 4 × 6.5
(800 ps) was assigned to the lifetime of guest molecules near Å2 for AEL-structures (MgAPO-11) (Figure 41A).390 Within
each other, and a 2.2 ns time constant was attributed to MgAPO-5 (AFI structure), H-aggregates are formed, and their
monomers that were not interacting with others.442 Another fluorescence is quenched strongly decreasing the quantum
example of a molecule that forms caged aggregates is Py, a yield: ϕ (%) = 0.4 versus 22 for PY in solutions (Figure
xanthene-derivative dye (Scheme 2). Different loadings of Py 41B).390,427 The larger pore dimensions of the AFI structure
encapsulated in zeolite L have been investigated.394 At highly allow the population of unrestricted face-to-face (H-aggregates)
diluted conditions (0.6% loading), the composite exhibits conformations, even at low concentrations.390,427 Their
similar behavior to that observed for the free Py in solution, population can be reduced by decreasing the Py concentration
indicating that the caged Py molecules are present as isolated and by increasing the Mg content.427 Inhibiting the formation
monomers.394 However, at higher concentrations (20% of this type of aggregates, by decreasing the pore size of the
loading), the caged Py molecules form J-aggregates, generating host, was reached using MgAPO-36 materials.390 In this case,
additional red-shifted transitions and shortening the fluores- the MgAPO-36 structure allows face-to-tail interactions and
cence lifetimes. At this loading, a biexponential fluorescence thus J-aggregate formation of Py (Figure 41B).
was observed, with lifetimes of 0.7 and 2.9 ns, assigned to J- Interestingly, Py monomers emit green (560 nm, 3 ns) light,
aggregates and monomers inside the zeolite, respectively.394 while J-aggregates in the composites emit red (660 nm, 1
Similar observations have been reported for perylene dyes ns).390 The interaction of PY with MgAPO-11 materials, whose
(N,N′-bis(2,6-dimethylphenyl)-3,4,9,10-perylenetetracarboxylic pore width (6.5 Å) is similar to the Py cross dimensions (6.8 Å,
diimide (DXP), N,N′-bis(3,5-dimethylphenyl)-3,4,9,10-peryle- measured perpendicular to its main axis), was investigated.
netetracarboxylic diimide (PR149), and N,N′-bis(1-hexylhep- Only the caged monomer was observed, and it emits with a
tyl)-3,4,9,10-perylenetetracarboxylic diimide (HPDI), Scheme longer lifetime than Py in solutions (3 vs 2 ns, respectively),
6) interacting with zeolite L.395 Interestingly, for these reflecting the rigidity that the matrix imposes on the monomer
composites, the presence of substituents on the pyrene movement, limiting then its nonradiative processes (Figure
structure can inhibit or favor the molecular packing.395 DXP 41B).390 Similar behavior was reported for rhodamine
exhibits aggregation, while PR149 and HPDI do not show it. dyes.391,392,424−426,443−451 When R6G (Scheme 5) intercalates
Another factor that obviously influences the formation of into solid thin films of laponite (Lap) clay, it tends to self-
aggregates is the concentration of the dye inside the zeolite aggregate, resulting in trapped H- and J-aggregates.426,447
cavities. This controllable factor provided to observe interesting Under very diluted conditions, only the R6G monomer was
behavior of DXP/zeolite L composites, whose spectroscopic observed, with an absorption band centered at 528 nm and an
properties strongly depend on the loading. Under diluted dye emission one at 548 nm.426,447 The average lifetime of the
suspensions, monomers emission is mainly recorded (with a monomer is 4.2 ns, which is slightly longer than in ethanol
lifetime of 3.8 ns), while at high dye loadings, a significantly solution (3.95 ns) and reflects the weak confinement effect on
shorter lifetime (from 1.7 to 0.7 ns), assigned to J-aggregates, the emission decays of the inorganic nanostructured materi-
was reported.395 In contrast, the voluminous groups of PR149 al.446,447 When R6G concentration increases, H- and J-
and HPDI prevent these dyes from packing at the sufficiently aggregates are formed.426,447 Their gradual formation is
short center-to-center distance that is needed for strong evidenced by the quenching of the fluorescence signal and
coupling.395 progressive decrease in the average emission lifetime from 4.2
Aggregates formation in SBMs was also controlled by ns (average lifetime in diluted solutions) to 0.45 ns (average
modulating the host properties. This has been demonstrated lifetime in concentrated solutions).447 These results show that
by encapsulating Py in aluminophosphate zeotype (APO) the ratio of H- to J-aggregates increases with the increased dye
AM DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 42. (A) Schematic representation of various sites at clay mineral surface, such as (a) external, (b) internal or interlayer, and (c) particle edges.
(B) Fluorescence decay curves recorded at 595 nm after excitation at 410 nm of rhodamine 3B/montmorillonites dispersions. Reprinted from ref
391. Copyright 2007 American Chemical Society.

loading.447 However, different behavior was observed in decay.266 When encapsulated in micro- and mesoporous
mesostructured 2d-hexagonal silica films.424 Steady-state experi- materials (NaX, NaY, MCM41, AlMCM41, and SBA15),
ments revealed that at low dye concentrations, H-aggregates HBA-4NP forms H- and J-aggregates. The absorption spectrum
preferably form, while at high concentrations, J-aggregates are of caged HBA-4NP has bands at 325, 370, and 410 nm,
favored due to the presence of the templating cylindrical assigned to the absorption of the H-aggregates, monomers, and
copolymer micelles that are highly organized into a 2D- J-aggregates, respectively. The emission spectrum exhibits a
hexagonal (p6mm) symmetry within the mesopores.424 The broad band with three intensity maxima (460, 480, and 530
aggregation of R6G has also been investigated in the presence nm), which are due to the emission of the K* forms of the
of organic−inorganic silica hybrids that were prepared using a monomer and J- and H-aggregates, respectively.264,265 The
sol−gel synthesis (bulk sample) and by spin-coating methods changes reflect the presence of these new species, whose
(thick film samples with a thickness ∼2 μm).451 For thick film formation is related to the size and composition of the host. For
samples, irrespective of the dye concentration, no aggregates example, when HBA-4NP is encapsulated in zeolites, different
were observed. However, in the bulk samples, J-type aggregates loading yields were observed, depending on the host (90% for
were formed, along with the corresponding decrease in the NaY while 20% for NaX).264 The characteristics of the space
fluorescence lifetimes, due to enhanced nonradiative tran- available within the host determines the loading yield; thus, in
sitions.451 Comparable behavior was reported when a similar NaY, which has fewer sodium cations and aluminum atoms in
dye, rhodamine 3B (Rh3B, Scheme 6), was adsorbed on the its framework, there is a stronger tendency to form dye
surface of montmorillonite layers (Figure 42A).391,392,445 The aggregates than in NaX.264 When HBA-4NP is trapped in
presence of diethyl substituents at the N atoms of Rh3B and of MCM41 and AlMCM41, whose identical pore sizes are wider
Li+ cations in the montmorillonite structure prevents the than that of the faujasite zeolites (2.5 vs 1 nm, respectively),
formation of H-aggregates and favors J-ones.391,445,452,391 The high loading yields (85 and 95% for MCM41 and AlMCM41,
high heterogeneity of the system results in four emission respectively) and larger amounts of aggregates were
lifetimes (∼40 ps and ∼ 250 ps and ∼ 1 ns and ∼2 ns) of the reported.265 However, there is less aggregates formation in
excited composites. When the system is heated, the population SBA15, a host with a wider pore size (∼6 nm), due to a weak
of J-aggregates increases, evidenced by the value of the average loading yield (40%).265 In all the studied complexes, the ESIPT
fluorescence lifetime, which changes from 710 ps at 0 °C to 460 reaction is very fast (<50 fs), leading to K* structures of the
ps at 130 °C (Figure 42B). The average lifetime of Rh3B in the monomer and aggregates. The fluorescence lifetime of the
hybrid complexes is shorter than that in water (1.53 ns), which caged K* monomer is remarkably longer in silica-based
indicates that an efficient quenching process occurs when Rh3B materials (∼6 ns in NaX and NaY and ∼3 ns in MCM41,
interacts with these materials.391 The decrease in lifetime values AlMCM41, and SBA15) than in dichloromethane solution (14
is explained in terms of aggregates formation that enables new ps).264,265 These changes in lifetime reflect the large degree of
deactivation processes such as internal conversion and homo restricted movement that is enforced by these hosts. The
EnT reactions. lifetimes of the trapped H- and J-aggregates are 100 ps and 1 ns,
Photochromic aromatic Schiff bases with an IHB, belonging respectively, and they become shorter with an increase in the
to the family of salicylideneanilines (SAs), also show a tendency guest concentration.264 The photobehavior reflects the
to form aggregates upon encapsulation.263−267 HBA-4NP, a presence of monomers and aggregates in the composites; the
derivative of SA, forms H- and J-aggregates when it is trapped formation and stability of the H- and J-aggregates depend on
in faujasite zeolites, MCM41 and SBA15 materials (Scheme the properties of the material. For example, SBA15 composites
3).264−267 In a pure dichloromethane solution, no aggregation have the lowest population of aggregates due to their low
was observed. The emission spectrum exhibits an intensity Brunauer−Emmett−Teller (BET) surface area (780 m2/g vs
maximum at 580 nm (Stokes shift ∼11,400 cm−1) due to the 825 m2/g for NaX, 1430 m2/g for NaY and 1000 m2/g for
formation of the K* tautomer that is created after an ultrafast MCM41/AlMCM41) and their surface roughness, which has
ESIPT reaction (<50 fs, see section 3.1) in the E* structure. fewer OH groups.265 In contrast, when MCM41 and
The K* species has a fluorescence lifetime of 14 ps, which is AlMCM41 composites with the same BET surface area and
very short because a twisting motion enhances the radiationless pore sizes are compared, the latter hosts more J-aggregates
AN DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 43. Representative emission transients of HBA-4NP in dichloromethane solutions and interacting with NaY cavities prepared from diluted (2
× 10−6 M) and concentrated (2 × 10−3 M) parent solutions. The insets show trapped monomers, H- and J-aggregates of HBA-4NP in the zeolite.
Reprinted with permission from ref 266. Copyright 2017 Elsevier B.V.

Figure 44. Cartoon showing the NR molecular structures (monomers, H- and J-aggregates) interacting with (A) X-MCM41 and (B) faujasite
zeolites. (A) Reprinted from ref 227. Copyright 2015 American Chemical Society. (B) Reprinted from ref 457. Copyright 2018 Elsevier B.V.

because of Al doping, which results in charge defects, increase for MCM41 and 450 fs and 4.5 ps for AlMCM41) than the
in the acidity of the pores, and thus the interaction between the dynamics in materials with smaller pores such as NaX and NaY
guest molecules and the framework. The host−guest and (300 fs and 3 ps).266,267
guest−guest interactions also play an important role in the NR also forms H- and J-aggregates in the presence of
ultrafast (fs-scale) behavior of these complexes. In dichloro- nanostructured host materials (Scheme 4). Its behavior in
methane solutions, the fluorescence of HBA-4NP exhibits two heterogeneous and organized media such as polymeric
ultrafast components that are assigned to IVR (100 fs) and a nanoparticles,453,454 micelles,318,455 cyclodextrins,317 zeo-
combination of VR/cooling and twisting motion (1.2 ps).266 lites,293,456,457 and proteins294 has been widely studied. It
These ultrafast processes slow down when HBA-4NP is forms aggregates in silica materials such as silica sol gels,
dissolved in a highly viscous medium (180 fs and 2.5 ps, laponite nanoclays, and MCM41 as a result of interactions with
respectively, in triacetin solutions), reflecting the internal the host matrix.226,227,458−461 In solution, it does not exhibit any
molecular motion in excited K*. This situation is reminiscent signs of aggregation.227,295,462,463 Time-resolved emission
of that of trapped HBA-4NP, as the confinement effect studies show that monoexponential decays (2.80 ns in
produced by the host, on aggregates formation, specific and methanol, 3.65 ns in ethanol, and 4.4 ns in dichloromethane)
nonspecific interactions with the host framework, leading to reflect the existence of only the monomer species.227,462 When
slower dynamics (Figure 43).266,267 Thus, in SBA15 complexes, NR interacts with laponite materials, it forms aggregates of
the IVR and VR/cooling processes have time constants (200 fs mostly the H-type.461 NR-laponite hybrid materials loaded with
and 2 ps) that are similar to those obtained in solutions, as the different amounts of the dye exhibit an unstructured broad
caged monomer species is predominant, and there is less absorption band that has two intensity maxima (550 and 600
restriction to movement.267 However, within NaX, NaY, nm) assigned to H-aggregates and the monomer species,
MCM41, and AlMCM41 materials, the values of these respectively.461 The decrease in the fluorescence quantum yield
components increase to 300−450 fs and ∼3−4.5 ps, from 34 to 14% in the composites of 0.06 and 0.5 NR
respectively, because of guest−guest interactions in the molecules per disk of the host material also demonstrates the
aggregates. The dynamics in the MCM41 and AlMCM41 presence of these aggregates that are not highly emissive.461
composites, materials with larger pore sizes and more OH The introduction of the cationic surfactant CTAB suppresses
groups on the framework, are slightly slower (370 fs and 3.4 ps the formation of H-aggregates in these hybrid systems and thus
AO DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

increases the fluorescence quantum yield in the laponite comparable, several differences could be detected. First, the
systems.461 In a series of MCM41 that was doped with different molecular packing is more efficient in the zeolite nanocages
metals (M = Al, Ga, Ti, and Zr), it was found that the despite their reduced dimensions (pore/cavity diameters of 35
composites fluorescence decay is multiexponential, in contrast and ∼13 Å for the mesoporous materials and the zeolites,
with the monoexponential behavior in pure solvent.227 When respectively). This is reflected in loading efficiencies higher for
NR interacts with regular MCM41, three lifetimes appear in the the zeolites (∼2.5 × 1018 NR molecules/g zeolite) than those
fluorescence decay (τ1 = 0.3 ns, τ2 = 0.9 ns, and τ3 = 2.5 ns). for the mesoporous materials (∼2.0 × 1018 NR molecules/g
The longest time was attributed to the NR monomers that were MCM41) under similar experimental conditions. These
attached to the surface of the host, while τ1 and τ2 were differences are due to electrostatic interactions between the
assigned to H- and J-aggregates, respectively, with the largest encapsulated NR molecules, the Na+ cations, and the
contribution arising from the J-aggregates.227 A significant aluminosilicate framework of the zeolite cavity, which facilitate
shortening in the fluorescence lifetimes was observed upon molecular trapping.266 While the populations of the caged
increasing the dye loading, supporting the explanation that the monomers and H-aggregates are similar, the one of J-aggregates
molecular packing of NR within the MCM41 nanopores leads is higher within the MCM41 mesoporous materials due to the
to the formation of H- and J-aggregates (Figure 44). higher surface acidity of these hosts, whose SiOH groups
The specific and nonspecific interactions of NR with metal- enhance H-bonding interactions with the guest molecules.227
doped AlMCM41 and GaMCM41 materials only slightly affect The increased rates of both the fast and ultrafast dynamics for
the fluorescence lifetimes and contributions of the NR species the excited NR/zeolite complexes compared to those in the
relatively to those in the NR/MCM41 system.227 This result NR/XMCM41 ones reflect stronger host−guest interactions in
was explained by the low amount of Al and Ga atoms in metal- the former composites. This effect is more pronounced for the
doped MCM41 (3% in AlMCM41 and 3.7% in GaMCM41). H-aggregates but is also observed in the case of the J-
However, the interaction of NR with TiMCM41 and aggregates, as well as for the ICT and VC processes. Within the
ZrMCM41 results in a decrease in the fluorescence lifetimes zeolite cavities, the spatial restriction strengthens the host−
of the monomers and H- and J-aggregates when compared with guest and guest−guest specific and nonspecific interactions, and
the behavior of NR/MCM41.293 The strong quenching as a result, the emission lifetimes of the involved species are
indicates the opening of a new decay channel that leads to shortened, while the ICT process occurs more efficiently.
faster decays of NR molecules when interacting with these Furthermore, the restricted motion also increases the vibronic
hosts containing transition metals (Ti and Zr). The interactions coupling between the S1 and S0 states, thus enhancing the
in these systems enable electron injection (EI, see section 3.2) nonradiative deactivation pathways.464
into trap states that are formed by the d-orbitals of the host In this section, we have shown how the encapsulation of
metal atoms (Lewis acidity).461 Furthermore, in a study that energy D and A molecules in SBMs enhances their stability,
used cometal-doped MCM41 materials (Ti-AlMCM41) with giving rise to more efficient FRET. The SBMs can act as spatial
varied Al and Ti contents, the H- and J-aggregate contributions separator between D and A molecules, preventing aggregation
were related to the content of Brønsted (Al doping) and Lewis and quenching processes, and as protective or insulating hosts
(Ti doping) acidic sites in the silica framework.226 An increase especially for semiconducting QDs or metal NPs. An important
in the Ti content induces a substantial increase in J-aggregates aspect to consider in EnT systems is the relative concentration
population and a shortening in the NR emission decays via ET of energy D and A molecules trapped in the SBMs pore. A high
to the titania sites.226 In contrast, the formation of J-aggregates concentration of the partners allows close contact between
is a result of localized and specific interactions between NR and them but also can give rise to a competitive self-quenching
Ti-atoms, reflecting the importance of the physicochemical process. Using 2D nanostructures and/or 3D nanoporous
properties of metal-doped SBMs for the generation and SBMs with larger pore sizes results in predominant formation
stabilization of these species. Composites of NR interacting of H-aggregates. However, in 3D materials with smaller pore
with faujasite (NaY)-type zeolites having different Na/Al ratios sizes and less metal doping in their structures, J-aggregates are
and charge balancing metals (Li+, Mg2+, and Cs+) in the favored species because of the limited free space and the
dichloromethane suspensions have been studied using steady- promotion of specific and nonspecific interactions with the
state and fs−ps time-resolved spectroscopy.457 The results framework. This kind of interactions determines the possibility
show formation of different populations, reflected in H- and J- of EnT and are related to the effectiveness of SBMs for the
aggregates, monomers, and surface adsorbed species (Figure development of new technological applications.
44B). The Na/Al ratio and the nature of the doping metal ion
affect the relative contribution of each population. The 4. SINGLE MOLECULE
bathochromic shift of both diffuse transmittance and emission The design of silica-based host materials is challenging and
spectra suggests strong interaction of the dye with Brønsted often presents difficulties in tuning and assessing the number,
and Lewis sites of the zeolites. The femtoseconds fluorescence distribution, and nature of acid sites. As we commented on in
transients show an ultrafast formation (∼200 fs) of a CS in NR the previous sections, along with these challenges, the
followed by a vibrational cooling (1−2 ps). The fluorescence distribution of the encapsulated guests and the changes in
lifetimes of the composites range from ∼100 ps to ∼2 ns, which their reactivity and photophysical properties remain particularly
decease at high Na/Al ratios. This behavior is explained in exigent issues. Single-molecule/single-nanostructure optical
terms of H-bond formation between NR molecules and the spectroscopy techniques are best suited to address these and
zeolite framework. Interestingly, the emission lifetimes change other open questions related to the characterization of hybrid
with the metal ion exchanged zeolites, due to the variation in complexes. The importance of these advanced tools to the field
the properties (size and polarization ability) of the exchange of materials chemistry in general and to SBMs in particular has
cation.457 Although the steady-state spectral properties of NR been highlighted by several recent reviews.17,48,465−470 The
interacting with the zeolites and MCM41 materials are application of single-molecule fluorescence methods to study
AP DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 7. Molecular Structures of the Compounds Interacting with SBMs and Discussed in the Single Molecule Section 4

mass transport in SBMs has been the topic of a number of cence spectroscopic methods have been employed to
reviews focused on studies that pursued deeper insights into investigate the microenvironmental acidity distributions in
certain defining properties of porous silica, such as the pore mesoporous aluminosilicate thin films that serve as models for
structure, host−guest interactions, and diffusion dynamics. solid acid catalysts. In the following, we will only comment on
These reviews provide examples from the literature that the recent contributions not discussed in previous re-
demonstrate how fluorescence microscopy is used for views.17,48,465−470
elucidating important aspects of porous silica materials and 4.1. Single Molecule Reactivity in Silica-Based Materials
heterogeneous catalysis (e.g., diffusion, reactivity, morphology, ((5′ and 6′)-Carboxy-10-(dimethylamino)-3-hydroxyspiro[7H-
intergrowth, accessibility, and catalyst deactivation).467,469 The benzo[c]xanthene-7,1′(3H)-isobenzofuran]-3′-one, C-SNARF-
experimental methods covered in these reviews include single- 1, Scheme 7) was used as an effective probe of the local pH in
particle tracking and confocal and wide-field fluorescence low-pH media (i.e., 1 < pH < 5).471 Histograms of the emission
microscopy. Another topic related to the development and intensity ratios of C-SNARF-1 single molecules were used to
characterization of SBMs in which single-molecule methods are elucidate the acidity distributions in mesoporous aluminosili-
of fundamental importance is heterogeneous catalysis. Few cate films with varying Al2O3 content. The data revealed an
reviews in the field of heterogeneous catalysis have demon- increase in film acidity with increasing Al2O3 content. The
strated the potential of these techniques for studying the results were interpreted in terms of a downward shift in the pH
reactivity of catalyst particles.48,465,466,468,470 Furthermore, of the local environment from pH ∼ 5 (pure silica) to pH ≤ 3
optical super-resolution imaging has recently been reviewed as the Al2O3 content of the film increased from 0% to 30%. The
in light of studies on heterogeneous and homogeneous distributions obtained are broader than expected for homoge-
chemical reactions at the single-molecule level, including neous samples, revealing a high degree of heterogeneity in the
reactions involving zeolites.17 The high spatial resolution of acidity properties of the studied films. Single-molecule
optical super-resolution imaging and its ability to monitor fluorescence-based imaging methods were used to study the
dynamic systems can reveal additional detailed information on catalytic activity of the zeolite ZSM-5 in cracking processes.472
single-molecule reaction/adsorption processes and single- The presence of Brønsted acid sites in the framework of ZSM-5
particle catalytic processes, including chemical kinetics and leads to the formation of fluorescent products, which were
reaction dynamics; active-site distributions on single-particle monitored with single-turnover sensitivity and high spatiotem-
surfaces; and the size-, shape-, and facet-dependent catalytic poral resolution. To obtain information on the number and
activities of individual nanocatalysts. Single-molecule fluores- distribution of Brønsted acid sites, and on the catalytic activity
AQ DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

of ZSM-5 materials, the acid oligomerization of furfuryl alcohol


(FFA) was used as a primary catalytic reaction. FFA forms
highly fluorescent oligomeric carbocations as result of its
interaction with Brønsted acid sites.472,473 The main conclusion
of that work was that the thermal treatment of ZSM-5
composites changes their reactivity: mild steaming at 500 °C
increases the surface reactivity, with local turnover rates up to 4
times higher than those of the parent ZSM-5 crystals, while
severe steaming at 700 °C significantly dealuminates the zeolite
material, leading to turnover rates that are 460 times lower.473
These differences were explained in terms of the aluminum
distribution, as severe steaming leads to the migration of Al- Figure 45. (A) Model for the localization of uncharged (in
species to extra-framework sites, blocking the pore accessi- hydrophobic core of the F127) and charged (into the hydrophilic
bility.473 In a more recent report, the acid-catalytic reactivity of corona region) probe in the F127 surfactant filling the nanometer-
ZSM-5 zeolite was demonstrated using a PT reaction sized cylindrical pores incorporated in mesoporous silica monoliths,
(oligomerization of styrene derivatives) as a probe.474 The (B) diffusion modes for anisotropic, and (C) apparent isotropic
oligomerization reaction gives rise to dimeric and trimeric diffusion. Reprinted from ref 475. Copyright 2015 American Chemical
species, which have different photostabilities and are sensitive Society.
to the solvent conditions. For example, when an apolar solvent
(n-heptane) is replaced by a polar one (1-butanol), the zeolite
reactivity decreases by several orders of magnitude due to the of NR molecules are located in nonpolar local environments
strong chemisorption of 1-butanol onto the Brønsted acid sites. with polarities similar to that of n-hexane.476 NR exhibits one-
Furthermore, the study concluded that the presence and dimensional diffusion, which is consistent with its confinement
stability of di- and trimeric species are also influenced by local in the cylindrical pores of surfactant-templated mesoporous
structural defects in the ZSM-5 framework. While the dimeric silica films. Single-molecule emission polarization measure-
species are located at the surface and in the middle of the ments have also demonstrated that NR molecules diffuse with
zeolite crystals, the trimeric ones, which tend to form their long axes aligned parallel to the long axis of the pores and
aggregates, lie predominantly on the outer surface.474 This that they are orientationally confined to ∼0.6 nm diameter
report indicates that the accessibility of the Brønsted acid sites pathways within the pores.476 The diffusion coefficient for 1D-
diffusing NR is also shown to be ∼103-fold lower than in bulk
can largely determine the local reaction rates and product
solution. These results were interpreted in terms of NR dyes
selectivity.
being confined in the hydrophobic cores of the micelles.476
4.2. Single Molecule Diffusion in Silica-Based Materials 4.3. Other Studies
Spectroscopic single-molecule tracking (sSMT) was used to Single-molecule spectroscopy was also used to study the rate of
directly visualize the diffusive motions of uncharged, cationic, surface-support degradation at silica surfaces upon exposure to
and anionic perylene diimide dyes (TPDI, SPDI, and OPDI, bases under a variety of deposition conditions.477 The kinetic
Scheme 7) in 1D nanochannels within flow-aligned meso- analysis of the obtained experimental results has shown the role
porous silica monoliths incorporating nanometer-sized cylin- of thermal annealing and the addition of blocking layers in
drical pores.475 The monolith pores were filled with Pluronic increasing stability, which is derived not from an increase in the
F127 surfactant, water, and alcohol. The study reports both stability of the surface linkage but from a decrease in the rate at
isotropic and anisotropic diffusion for all three dyes. The which base molecules approach the surface before irreversible
charged dyes exhibit predominantly isotropic motions, attack at the silyl ether linkage. Furthermore, it was suggested
suggesting that these molecules readily pass through defects that the structural diversity of the silica surface is likely to be the
in the silica pore walls. In contrast, the motions of the cause of the observed heterogeneity in the surface-linkage
uncharged PDI became more anisotropic as the monolith ages. disruption kinetics.477
The results obtained from flow-aligned F127 gels in the In similarity with other nanostructures, SiO2 nanoparticles
absence of silica demonstrate that partitioning plays an can possess luminescent defects in their regular structure. A
important role in limiting the passage of the PDIs between photoluminescence study on defect centers in single 11 nm
pores. The fluorescence correlation spectroscopy (FCS) data SiO2 nanoparticles has shown that the photophysical properties
confirmed the presence of two populations of diffusing of these luminescent centers are comparable to those of single
molecules, attributable to anisotropic and isotropic motions. dye molecules.478 It was observed that the fluorescence from
All three dyes exhibit mean diffusion coefficients that decrease these particles exhibits a linear transition dipole and symmetric
with monolith age. On the basis of combining the sSMT and emission and excitation spectra. However, these luminescent
FCS data, aging-time-dependent chemical changes to the defects have presented larger variability from particle to particle
materials were suggested, which lead to partitioning of the due to variations in the local chemical environment around
charged PDI dyes into more polar regions that facilitate their each center of each particle.478 In a consecutive single particle
passage through the pore walls. In contrast, the uncharged PDI study, the photoactivation of the luminescent centers in SiO2
become better confined to the hydrophobic micelle cores with nanoparticles with sizes between 11 and 166 nm was
monolith aging, leading to enhanced 1D diffusion (Figure demonstrated.479 The work has shown that by exposing a
45).475 single SiO2 nanoparticle to UV illumination, it is possible to
However, this report did not resolve the location of the dye. create new luminescent centers within these particles. Due to
Thus, sSMT studies using the polarity-sensitive dye NR the much weaker chemical bonds in the SiO2 particles, it is
(Scheme 4) have been performed, showing that the majority possible to generate new defects in the nanostructures using
AR DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

UV light, which allows for the reactivation of the nanoparticle’s variation in its optical properties was also recorded, as revealed
fluorescence after its photobleaching.479 by steady-state and time-resolved emission techniques.484 The
4.4. Electronic Nanoconfinement study did not report any effect of the confinement of the
nanochannels on the ESIPT reaction (section 3.1) in the E*
As we comented on in the above parts, the encapsulation of
guest molecules in the channels of micro- and mesoporous structure of HBTNH2 to give K* tautomer, probably as it is
hosts may significantly change the electronic properties of the very fast (<100 fs) and it is not expected to be affected.
guest species, which may give rise to variations in optical band However, the component assigned to the excited-state lifetime
gaps and in the excited-state lifetimes of the organic molecules of the twisted K* of HBTNH2 increases from 22 ps in
in these composites. When a tight host/guest fit is observed, acetonitrile to 40−45 ps within MCM41 in support for the
these changes are reflected in variations in basicity and the electronic confinement effect on HBTNH2.484 Experimental
oxidation potential, for example. These apparent variations evidence for the electronic nanoconfinement of HBTNH2 at
cannot be explained only in terms of the influence of the host the single-molecule level has also been reported.210 Time- and
electrostatic fields on the stabilization of polar guests. An spectrally-resolved studies at the single-molecule level have
electronic confinement concept has been proposed based on demonstrated that depending on the way of the guest−host
the idea that the molecular orbitals (MOs) of the guests inside complex preparation, different sites of the host (MCM41) are
the host (zeolite) pores are not extended throughout the space accessed by the guest.210 The averaged emission spectrum of
but instead are limited to within the zeolite walls.480,481 This the single-molecule complexes (HBTNH2/AlMCM41) reveals
confinement effect is stronger as the size of the confined guest a single band centered at 583 nm with a full-width at half-
approaches the zeolite cage dimension, producing an energy
maximum (fwhm) intensity of 64 nm (∼1480 cm−1). The
increase in all MOs, particularly those that are more diffuse.
Theoretical studies have demonstrated that confining organic emission is assigned to a trapped anion (deprotonated phenol
molecules in the pores and cavities of SBMs is sufficient to alter moiety) of the dye. The spectrum is much narrower than the
their electronic properties as a result of changes in the one of K* generated using the undoped MCM41 (∼3250
molecular orbital energies and band gaps.480−482 Results from cm−1) samples (HBTNH2/MCM41 (Figure 46).
the calculations of the electronic properties of confined
benzene, naphthalene, and anthracene have been presented in
support of the electronic confinement effect.482 These
calculations have demonstrated that the HOMO is more
sensitive to confinement than the LUMO, and the overall effect
is a reduction of the band gap of the frontier MOs when the
guest−surface distance is less than ca. 2.5 Å. In that report, the
confining space of the cavities has been modeled using a mica
sheet with a molecule−surface distance in the range of 1.5−4.0
Å. Evidence of confinement has been revealed by semiempirical
calculations, which were interpreted by means of the Hückel
molecular orbitals theory. More recent calculations predicting
redshift of both absorption and emission bands of the dyes
upon encapsulation are commented in section 2.2. An
experimental evidence of electronic confinement was reported
for anthracene encapsulated within zeolites.464 The study has
reported a bathochromic shift of the 0−0 transition from 375
nm in 3-methylbutane to 396 nm in NaY and to 419 nm in
ZSM-5, showing that the smaller the pore size of the zeolite, the
larger the bathochromic shift of the 0−0 transition. The
observed shift along with the shortening in the fluorescence
lifetime (from 4.5 ns for free anthracene to 3.2 ns within NaY
zeolite and <100 ps in mordenite or ZSM-5) have been
correlated with the increase in the HOMO energy and with the
reduction in the HOMO−LUMO band gap. Other studies have
provided additional experimental evidence for the electronic
nanoconfinement effect in MCM41 and zeolite materials. For Figure 46. Single-molecule microscopy results of HBTNH2 covalently
example, significant variations in the photophysical properties bonded to silica nanoparticle and AlMCM41. (A) Emission decays of
of Py molecules incorporated inside zeolite crystals that have HBTNH2 covalently bonded to silica and (B) AlMCM41, along with
similar channel sizes, such as zeolite L and AlPO4−5, were the best nonlinear least-squares monoexponential fits (solid lines). (C)
observed by means of fluorescence polarization.483 The results and (D) are the distribution of the lifetimes from the monoexponential
indicate that the dye molecules’ optical properties are strongly fit to the decays of single-molecule silica and AlMCM41 composites,
respectively. (E) and (F) represent the average emission spectrum of a
affected by the framework structure of the host, especially the
single HBTNH2 covalently bonded to a silica nanoparticle and
specific geometry of the zeolite channels, which induces an AlMCM41 nanochannels, respectively. The dashed line in part (E)
angle dependence of the guest fluorescence behavior. The represents the Gaussian fit to the main contributing band in the
change was explained in terms of the electronic confinement average spectrum of single HBTNH2-silica complexes in the presence
effect. When the ESIPT dye HBTNH2 (Scheme 3) is of the strong base 1,8-diazabicyclo[5.4.0]undec-7-ene. Reprinted with
encapsulated within MCM41 and Al-doped MCM41, a strong permission from ref 210. Copyright 2010 American Chemical Society.

AS DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

This difference, along with the narrower lifetime distribution, lifetime, and quantum yield.478 Both the emission and
is a clear evidence of a stronger confinement of the local excitation spectra of different luminescent centers have shown
nanoenvironment around the single anion of HBTNH2 in the a shift within the range of 2.0−2.4 eV. Despite this broad
AlMCM41 channels. It is suggested that the replacement of Si variation, the shape of the spectrum remains constant, which
atoms with Al ones changes the electronic density distribution suggests that the fluorescence comes from the same type of
within the framework of MCM41, which in turn affects the defect and is a demonstration of the electronic nanoconfine-
host−guest orbital interaction that the resulting emission ment effect on the properties of the luminescent defects in the
comes from a much narrower density of the states of the SiO2 nanoparticles.478
single chromophore. Studying the single-molecule behavior of It has been shown that the rigid frameworks of zeolites can
DY-630-MI, a hemicyanine dye (Scheme 7), showed that act as a host in which silver clusters can be stabilized, enabling a
replacing a few Si4+ ions with Al3+ ones in the regular MCM41 high degree of control over their optoelectronic properties
changes the local electrostatic field within the nanotube, which (Figure 48).486,487 The dependence of their electronic proper-
results in a more selective orientation of the DY-630-MI ties on spatial confinement was studied by characterizing the
molecules.485 Comparing the distribution of the polarization ionization potential of the clusters embedded in four different
(P) values for DY-630-MI/AlMCM41 with those for the free zeolite environments over a range of silver concentrations.487
molecule and the DY-630-MI/MCM41 complexes, a stronger The chosen faujasite frameworks (FAUX and FAUY) differ in
bias (61%) toward the positive limit (P = 0.71) is observed the Si/Al ratio of the framework, while the two LTA zeolites
(Figure 47). This behavior is indicative of an increase in the (3A and 4A, Figure 49A) differ in the alkali metal charge-
balancing cation (Na+ and K+, respectively). The results reveal a
strong influence of silver loading and of the host environment
on the Ag cluster ionization potential (over a range of >0.5 eV),
which is also correlated with the cluster’s optical and structural
properties (Figure 49). Using zeolites as host, the photo-
luminescence quantum yields of the composites could be
optimized to reach nearly 100% by careful tuning of the
mobilities of nonframework metal cations.487
Single-molecule fluorescence and single particle tracking
methods reveal details of mass transport, catalytic reactions, and
molecular distribution in SBMs at the single-nanostructure and
single-molecule levels, while also providing information on their
Figure 47. (A and B) Emission polarization (P) distribution physical morphology. Additional information on molecular
histograms for DY-630-MI interacting with MCM41 and with diffusion coefficients and degree of molecular confinement can
AlMCM41 materials, respectively. Reprinted with permission from be obtained by using single molecule spectroscopy.
ref 485. Copyright 2011 Royal Society of Chemistry.

5. RECENT APPLICATIONS OF SILICA-BASED


number of more oriented single chromophores in the MATERIALS
nanochannel due to the modified electrostatic environment
around the encapsulated single molecule.485 5.1. Photocatalysis
In a study of the photoluminescence of defect centers in Micro- and mesoporous SBMs are widely used in the field of
single 11 nm SiO2 nanoparticles, it was suggested that a catalysis.3,13,22,23,488 This is mainly due to the special and
strongly inhomogeneous local chemical environment around interesting features of these materials such as their high surface
the centers results in a broad and random variation of the single area, high thermal and hydrothermal stability, good adsorption
nanoparticle emission and excitation spectra, fluorescence capacity, and the presence of actives sites, commonly Brønsted

Figure 48. Silver nanoclusters encapsulated in LTA and FAU sodalite cages, which are connected via the secondary building units (double 4-ring for
LTA and double 6-ring for FAU). Reprinted with permission from ref 486. Copyright 2016 Macmillan Publishers Limited, part of Springer Nature.

AT DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 49. (A) Variation of the ionization potentials of the four calcined zeolites (FAUX[Ag11], FAUY [Ag6.5], 3A[Ag12], and 4A[Ag12]) loaded with
different contents of silver. (B) Peak emission energy change with the ionization potential for the studied zeolites. Reprinted with permission from
ref 487. Copyright 2016 Macmillan Publishers Limited, part of Springer Nature.

and Lewis acid ones.10,488 Furthermore, due to the dimensions process induces formation of •OH radicals, which leads to the
of their channels, these materials can also be selective toward activation and acceleration of norfloxacin photodegradation.
reactants, products, and/or transition states; as such, they can The presence of Ag enhances the visible light absorbance and
be used for direct catalytic reactions while avoiding undesired provides better charge separation. Moreover, these composites
side processes (see Scheme 8 for molecules discussed in section show high stability even after being used several times.500
5). Acid and base catalytic reactions are used at a large scale in Another photocatalyst formed by Ag was studied for the
the petrochemical industry for isomerization, hydrocarbon photocatalytic degradation of MB (Scheme 4).501 In this
cracking, and hydrocarbon synthesis.26,77,488−492 The catalytic system, the MB interacts with thiol-functionalized MCM41
redox process generally occurs due to the presence of metal nanofibers (MCM41-SHNFs) in the presence of silver bromide
atoms in the framework and results in the oxidation of the (AgBr).501 The photocatalytic reaction was observed at 664 nm
reactant.22,26 Numerous reviews have reported on the use of to follow the photodegradation and photodecolorization of
zeolites and mesoporous materials in catalytic reac- MB. In the absence of photocatalyst, the degradation of MB is
tions.3,10,15,79,439−441 Therefore, in this review, we will highlight negligible. However, in the presence of AgBr/MCM41-SHNFs
only the most recent reports, predominantly those using light. composites and after 60 min under visible light irradiation, 87%
Photocatalytic reactions are initiated by light. In these of MB is removed (Figure 50A).501 In the presence of AgBr
reactions, the encapsulated reactant is usually a photoactive alone, the photodegradation process occurs with a lower
organic or inorganic molecule, although it can also be produced efficiency. Only 35% removal of MB after 60 min was observed.
by the presence of heteroatoms (Ti, V, and other transition This indicates that MCM41-SHNFs as a support plays an
metals) in the micro- and mesoporous frameworks. Photo- important role in this photocatalytic reaction due to its ability
catalytic reactions are present in a wide variety of processes to adsorb more MB molecules on its surface, has more reactive
such as the photoreduction of CO2 by H2O, the photocatalytic sites, generating larger amounts of free radicals during the
oxidation of saturated hydrocarbons, the photosplitting of water reaction, which contributes to the degradation of the MB
into hydrogen and oxygen molecules, and the photogeneration molecules. Furthermore, its presence prevents the photo-
of hydrogen peroxide, among others.18,493−498 All such corrosion of the AgBr semiconductor, granting an 87% removal
reactions are of great importance for the environment and of MB even after five successive cycles (Figure 50B).501 AgBr/
industrial chemical processes for the development of renewable MCM41-SHNFs composites showed higher stability with
energy alternatives to fossil fuels. respect to AgBr and other composites such as AgBr/SBA15,
In the presence of semiconductors and/or metallic atoms, AgBr/ZnO, and AgBr/TiO2.501,506,507
the SBMs can take part in photocatalytic degradation of the This reaction was also investigated using other mesoporous
encapsulated dye.499−505 For example, NaX zeolite impregnated materials, such as SBA15 containing TiO2 semiconductor
with Ag/AgCl−zerovalent iron particles (ZVIP) was prepared nanoparticles.502,503 The results showed an enhanced photo-
for tetracycline (TC, Scheme 8) degradation.499 The reaction in catalytic efficiency of MB decomposition, exhibiting rapid total
aqueous solutions under visible light is negligible due to the low decolorization. Furthermore, doping SBA15 with Al atoms
TC absorption in the visible range. However, the photocatalytic (AlSBA15) creates a decolorization efficiency even higher than
efficiency of the composites is enhanced with the increase in those of TiO2/SBA15 and commercial TiO2. This was mainly
the ZVIP particles contents in the zeolite structure, displaying attributed to a more effective adsorption capability resulting
maximum TC degradation when its weight ratio is 5%. Thus, from the increased specific surface area after the substitution of
the TC and ZVIP interaction allows the formation of Si species with Al ones.503 Another recent report showed the
composites with a strong absorption in the visible light region efficiency of a zeolite composite, graphene-Bi8La10O27-zeolite
and greater charge separation efficiency.499 Another report has (GBZ), for the degradation of MB and other dyes such as RhB
shown the degradation of norfloxacin in zeolite containing (Scheme 5) and Texbrite (TXB).504 For this composite, the
silver oxide (Ag2O) and decorated with TiO2 (Ag2O/TiO2- typical absorption peaks at 555, 665, and 360 nm for RhB, MB,
zeolite).500 Norfloxacin is an important antimicrobial drug but and TXB, respectively, decrease by 85, 87, and 83% after 5 min
with a potential risk to human beings and the environment. For of visible light irradiation. Their total degradation was observed
this reason, its photodecomposition has been evaluated in these after 10 min. Upon visible light irradiation, high energy-
systems. In Ag2O/TiO2-zeolite composites, the photocatalytic photoelectrons are detected in the zeolite composites, whose
AU DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 8. Molecular Structures of the Compounds Interacting with SBMs and Discussed in the Applications Section 5

AV DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Scheme 8. continued

Figure 50. (A) Changes in the UV−visible absorption spectra of MB showing the photodegradation under visible light irradiation in the presence of
AgBr/MCM41 SHNFs nanocomposites. (B) Cycling degradation efficiency of MB solution in the presence of AgBr nanoparticles and AgBr/
MCM41 SHNFs nanocomposites. Reprinted with permission from ref 501. Copyright 2016 Elsevier B.V.

Figure 51. Schematic representation of graphene-Bi8La10O27-zeolite nanocomposite, showing a high photocatalytic efficiency activity toward organic
dyes degradation. Reprinted with permission from ref 504. Copyright 2017 Elsevier B.V.

formation leads to the generation of •OH radicals that are irradiation were necessary to cause the almost total degradation
responsible for the photodegradation of the free dyes (Figure (∼98%) of rhodamine. The photocatalytic process is initiated
51).504 The photodegradation of rhodamine molecules in when TiO2 absorbs the UV radiation, generating electron−hole
aqueous solutions with TiO2-chabazite materials was also pairs on the zeolite composites. The high oxidation potential of
investigated.505 For these composites, 90 min of light the hole in the catalytic surface and the presence of •OH
AW DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

radicals cause the dye photodegradation but with an efficiency capacity of the catalyst for PNT. Thus, an optimal loading of 2
lower than in the case of GBZ.505 This is explained in terms of w/w % is necessary for these composites to reach their
higher molecular adsorption capacity of GBZ, as this system maximum efficiency.516 This efficiency was improved with the
contains two different surfaces (graphene-Bi8La10O27 and copresence of Ag ions in the zeolite framework, as Ag particles
zeolite) for the adsorption of the molecule, which can absorb help to generate a higher charge separation and thus improve
more light and generate more electron−hole pairs to remove the photocatalytic activity of the catalysts. This was also
the dye pollutant molecules. observed in other systems used for the photoreduction of water
The use of micro- and mesoporous composites containing to hydrogen and oxygen molecules.517 The coupling of the
TiO 2 presents remarkable advantages due to its easy TiO2 semiconductor nanoparticles present in zeolite Y with a
preparation, low cost, and chemical stability.508−511 A recent metal (Pt) and another semiconductor (CdS) has a stronger
study reported on the adsorption capacity of TiO2 immobilized influence on the photocatalytic activity of the formed
in a zeolite framework for humic acid detection and extraction composites. The hydrogen evolution rate improves by a factor
from water.512 In the absence of TiO2, the natural zeolite of 100 and 18 when Pt and CdS were incorporated to the TiO2-
showed a low affinity for humic acid, while in its presence, the zeolite Y system, respectively.517 It is worth noting that to
formed composite was an effective adsorbent that removed the obtain these results, the titanium, which is part of the zeolite
humic acid fraction with high aromaticity and molecular weight. framework, has to be in intimate contact with Pt or CdS. For
Furthermore, the TiO2/zeolite composites are readily regen- this reason, in the Pt/CdS/TiO2-zeolite Y ternary system,
erated by UV light irradiation, so that the adsorption capacity is where the localization of nanoparticles is mediated by the
maintained even after five successive adsorption−regeneration zeolite via a self-assembly process, the improvements in the H2
cycles.512 Another report presented the photodegradation evolution rate were not as marked because Pt was not in direct
reaction of acetaldehyde gas in the presence of TiO2/zeolite contact with TiO2.517 Another system composed of titania and
type X (Figure 52).513 The results demonstrated that the iron-exchanged zeolite was used for the photodegradation of
photocatalytic decomposition activity in TiO2/zeolite X was diclofenac (DCF, Scheme 8).518 This anti-inflammatory drug is
higher than in TiO2 nanoparticles.513 a pollutant that, when present in water, adversely affects human
health and the environment. DCF adsorption onto TiO2−FeZ
surfaces and the subsequent irradiation leads to its photo-
degradation.518 However, the generated photoproduct tends to
accumulate on the TiO2−FeZ surface, thus negatively affecting
the subsequent photocatalytic cycles. Therefore, thermal and
chemical reactivation strategies were necessary to improve the
photocatalyst stability in reuse cycles of these systems. Thermal
treatment at 553 K produces an increase in the activity of
consecutive cycles due to a high DCF and total organic carbon
(TOC) removal rate and an increase in TiO2 crystallinity. A
chemical treatment using ozone flow also increased the reuse of
this material, although with an activity lower than in the
previous treatment.518
The combination of SBMs and some organic or inorganic
Figure 52. Schematic representation of the photodegradation reaction compounds also gives rise to formation of active photocatalytic
of acetaldehyde gas in the presence of TiO2/zeolite X. Reprinted with
permission from ref 513. Copyright 2013 Springer.
systems.519−521 For example, tungstosilicic acid (TSA, Scheme
8) immobilized on NH4Y and NH4ZSM5 zeolites was used for
the photodecomposition of methyl orange (MeO, Scheme
A similar result was observed for the photodegradation of 8).519 Heteropolyacids (like TSA) were used as photocatalysts
methanol gas using TiO2/β-zeolite and TiO2/SBA15 materials, in the degradation of organic pollutants present in water due to
where the reaction efficiency was higher than in conventional their nontoxicity and high photostability. These compounds
P25-TiO2 commercial nanoparticles.514,515 The result was showed photocatalytic properties comparable to semiconduc-
explained in terms of a higher photocatalytic activity of TiO2 tors such as TiO2, ZnO, and CdS. However, these systems
when it is supported, as the micro- and mesoporous structures cannot be used directly in catalytic reactions because they have
improve the separation of the electron−hole charges formed in low surface area and high solubility in polar solvents, which
the semiconductor under irradiation. Furthermore, the acidity makes them impossible to reuse. When supported in zeolite
and the presence of active sites on the surface of these materials materials, the photocatalytic system increases its specific surface
lead to a high adsorption of organic compounds and an increase area and allows their easy recovery and reuse.519 The amount of
in their photoreactivity. Another report showed the photo- TSA (% w/w) on the zeolite surface also affects the efficiency
catalytic activity of zeolite Y impregnated with different of the photocatalytic reaction. After 240 min with 5% w/w
amounts of TiO2 in the degradation of p-nitrotoluene (PNT, TSA, the photocatalytic system removes 28% of MeO, while
Scheme 8) in aqueous media.516 The degradation of PNT with 30% w/w and in the same time range, the photo-
adsorbed in these composites gives rise to a mineralization degradation process was improved, reaching 75% removal.519
resulting in simple final products: CO2, H2O, NO3−, and NH4+. However, despite being supported, this system showed a low
The activity of this photocatalytic process depends on the reuse capacity, with the catalytic efficiency decreasing from 75%
amount of TiO2 in the zeolite structure. A low loading of TiO2 to 36% in the second cycle. This is due to the adsorption of the
leads to a lower generation of charge-separated states and reaction products on the catalyst surface, decreasing the
therefore less photodegradation, while a high loading can block number of actives sites available for the absorption of the
the zeolite pores, resulting in a decrease in the adsorption initial reactant.519 Another composite formed by cobalt
AX DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 53. (A) Fluorescence spectra of RhB in tetraphenylethene-bridged periodic mesoporous organosilica particles at different dye contents. (B)
CIE chromaticity coordinates showing the emission colors of the films (dots). Reprinted with permission from ref 525. Copyright 2015 Royal
Society of Chemistry.

phthalocyanine (CoPc, Scheme 8) immobilized onto MCM41 systems can show collective EnT and ET with possible
was used for the degradation of MeO.520 After 2 h of visible functions such as signal amplification and light harvesting.
light irradiation, the photodegradation process was completed, More recently, mesoporous silica nanoparticles of 50 nm in
with a removal rate of 99%. This photocatalytic efficiency is diameter were prepared containing two chromophores that
much better than that observed in the previous study and in emit in different spectral ranges.524 One of them was the Zn
another system formed by a semiconductor, CuO, incorporated complex of HBT (Scheme 3), emitting blue light, while another
in zeolite X, where only 60% of the dye was removed under the was the orange-red dye RhB (Scheme 5). By choosing the
same experimental conditions.522 Furthermore, CoPc/MCM41 concentrations and ratio of these two fluorescent centers, a
systems exhibited high photostability, being useable for four nearly white light-emitting system was produced.
consecutive cycles, with the MeO degradation process only The coupling of two chromophores of opposite fluorescence
decreasing from 99 to 90%.520 These results revealed that the behaviors was also realized in periodic mesoporous organo-
use of SBMs as support and reaction hosts provides high silicas.525 Blue-emitting tetraphenylethene-bridged organosilane
stabilization for the organic and inorganic guests in addition to was used as a precursor for preparing high-performance
preventing their molecular aggregation, so improving their luminescent nanoparticles. These NPs avoid the typical
catalytic activity. fluorescence quenching and thus are aggregation-induced
5.2. Photonics emission materials. RhB (which exhibits aggregation-induced
quenching) was then encapsulated in the nanoparticle as an
One of the most important issues in photonics field is the study energy acceptor partner. The emission of the resulting
and development of light sources. The application of silica- composites could be fine-tuned over the entire visible spectrum
based hosts in designing efficient, broadband, and stable light by adjusting the content of the encapsulated RhB dye, as
sources has great potential. First, as we already commented and illustrated in Figure 53. Highly pure solid-state white light was
highlighted, encapsulation of organic and inorganic fluoro- achieved with a quantum yield of nearly 50%.
phores can enhance their emission as a result of suppression of In another contribution, the aggregation-induced emission
the nonradiative deactivation processes (see sections 2 and 3). molecule coded CWQ-11 (Scheme 8), a carbazole-based
Furthermore, as commented on in this review silica hosts cyanine molecule, has been doped in situ into SiO2 NPs.526
enable the coloading of chromophores emitting in different The fluorescence of such molecules is closely related to their
spectral regions, facilitating broadband lighting sources. Finally, restricted structure. Inorganic silica NPs provided a rigid
encapsulation usually provides protection against the outside microenvironment that stabilized CWQ-11 molecules to obtain
environment, which is important for improving the stability of fluorescent materials exhibiting almost 50 times higher
the light source. A few recent examples of the above strategies quantum yield than that for the same molecule in solution.
will be outlined below. The examples are grouped into four Both one- and two-photon fluorescence was studied for
categories, depending on the nature of the guest species: hemicyanine molecules encapsulated in nanosized silicalite-1
molecules, rare earth elements, silver clusters, and perovskites. zeolite.527 It was found that one- and two-photon absorption
The studies of dye-doped silica nanoparticles before 2010 coefficients were the same for the dye in solution and inside the
were reviewed in another report.373 A system with ruthenium zeolite. However, significant differences were found in the
tris-bipyridyl encapsulated inside zeolite Y supercages was fluorescence quantum yields where the one-photon fluores-
tested for use in electrochemiluminescent cells.523 The cence was enhanced 5 times, while the two-photon signal was
synthesis of the ruthenium tris-bipyridyl complex was achieved enhanced 3.3 times upon encapsulation.
by a ship-in-a-bottle method. The cell operated at an optimum Lanthanide ions possess unique and attractive optical
voltage of 3 V. The electrochemiluminescent efficiency of the properties such as sharp and long-lived emission in the visible
cell was significantly increased upon the ion exchange of to near-infrared (NIR) region with high color purity. However,
sodium with cesium and the vapor deposition of calcium metal their practical application is hindered by their poor physical
inside the zeolite pores. Generally, in such configurations, each stability, which can be overcome by immobilization in SBMs.
nanoparticle can contain several active molecules, allowing very The luminescence properties of zeolite functionalized with
high absorption coefficients (>106 L mol−1 cm−1) and lanthanides were recently reviewed.528 In addition to improved
luminescence quantum yields. Moreover, multichromophoric thermal stability and photostability upon encapsulation, other
AY DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 54. (A) Emission of Eu3+ in zeolite supercages induced by UV irradiation upon washing with the indicated solvents and calcined. Reprinted
with permission from ref 528. Copyright 2014 Royal Society of Chemistry. (B) Electroluminescence spectra of ZEOLED (silver-exchanged zeolites)
at different applied voltages. Reprinted with permission from ref 530. Copyright 2017 Wiley-VCH.

Figure 55. (A) Emission decays of CsPbBr3 nanocrystals synthesized within meso-SiO2 of different characteristic size. The inset shows the sizes of
the crystals. Reprinted from ref 533. Copyright 2016 American Chemical Society. (B) Normalized emission spectra of MAPbI3 nanocrystals grown
using templated TiO2 (orange) and SiO2 (purple) porous films as scaffolds. Gray dashed spectrum corresponds to the photoemission of a flat
perovskite film. Reprinted with permission from ref 537. Copyright 2017 Wiley-VCH.

new properties of lanthanides were also reported, including second emission band at 545 nm with a much longer lifetime
spectrally tunable luminescence and emission enhancement (microsecond time scale) was attributed to charged silver
upon annealing (Figure 54A). Moreover, multicolor-emitting species. Introducing luminescent silver-exchanged zeolites into
lanthanide complex-loaded zeolite materials could be fabricated, a conductive polymer matrix as emitters resulted in new
and transparent luminescent composites could be created by electroluminescence bands.530 The successfully constructed
embedding nanosized Ln3+‑functionalized zeolites into poly- devices were called ZEOLEDs. Their electroluminescent
(methyl methacrylate) (PMMA) matrices. Even more complex spectra showed additional emission bands besides those present
hybrids were proposed based on incorporating a rare earth in photoluminescence (Figure 54B). The emission color of this
ternary complex into CdS QDs-loaded zeolite Y through a system [Commission Internationale de l’Éclairage (CIE)
coordination reaction.529 The highest luminescence quantum coordinates] could be tuned from blue to red color by
yield was obtained for the system consisting of CdS−zeolite changing the silver concentration in the zeolite framework.
functionalized with europium, mercaptoacetic acid, and 1,10- Perovskites emerged as new and very promising materials for
phenanthroline ternary complex, with a lifetime of almost 1 ms. photovoltaics. A few examples of combining SBMs with
An interesting alternative to the above luminescence perovskites for solar cells application will be presented in the
materials was recently proposed based on silver clusters next section 5.3. Simultaneously, the slow charge recombina-
incorporated in a zeolite matrix.531 Direct evidence at an tion and excellent charge transport properties of perovskites are
atomic level of luminescent silver species stabilized in faujasite also very promising in luminescence applications. A few
zeolites was presented using scanning transmission electron strategies to employ silica templates to improve such
microscopy (STEM). Two different silver clusters were luminescence properties will be outlined below.
identified: a trinuclear silver species associated with the green In one of the first such reports, a large variety of perovskite
emission, and a tetranuclear silver species associated with the compounds, hybrid and fully inorganic, were studied.533 With
yellow one. A nonluminescent state associated with Ag3 species the general perovskite formula APbX3, the materials were tested
was found to be a precursor for the subsequent formation of in which A was cesium (Cs), methylammonium (MA), or
luminescent Ag clusters upon heat treatment. Stable formamidinium (FA), and X was Cl, Br, I, or a combination
luminescent silver clusters were also observed in sodalite between two of them. Various commercially available
cages of hexagonal analog faujasite type zeolite (EMT).532 The mesoporous silica templates of pore sizes ranging from 2.5 to
octamer (Ag8) was proposed as one of the most stable clusters 50 nm were used to encapsulate the zeolite QDs. These were
and exhibited molecular-like emission properties with an synthesized inside the pores by infiltration of the precursor
emission band at 395 nm and a nanosecond decay time. A solutions, and then after drying, template-assisted perovskite
AZ DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 56. (A) SEM images of zeolite L monolayers, and zoomed region of the composite containing two dyes and modified with a stopcock.
Reprinted with permission from ref 388. Copyright 2008 Royal Society of Chemistry. (B) Schematic representation of energy transfer (EnT) process
in a chromophore embedded in periodic mesoporous organosilica. Reprinted with permission from ref 542. Copyright 2014 Royal Society of
Chemistry.

nanocrystals were formed. This simple method led to a high that perovskite nanocrystals can become stable visible light
emission quantum efficiency exceeding 50%. In Figure 55A, an emitting materials at a desired spectral region.
example of testing trapped CsPbBr3 crystal in SBMs is shown, Another branch of photonics with interesting applications of
for which the longest perovskite lifetime (the most luminescent SBMs is in light transmission media. The most explored silica-
composite) was observed in the silica material of 7 nm pores based material in this context is probably zeolite L. Its crystals
(hexagonally ordered 1D-channels known as MSUH meso- can contain oriented fluorophores in their parallel nano-
phase). channels. They then possess remarkable fluorescent properties
Green-emitting CsPbBr3 crystals were also mixed with and can be arranged in nearly any desired manner via self-
mesoporous silica of pore sizes 12−15 nm to prevent the organization methods.388 FRET between encapsulated dyes is
anion-exchange effect.534 The composite was then mixed with the mechanism responsible for energy transport along the
red CsPb(Br0.4I0.6)3 nanocrystals in silicone resin and then channels. Communication between the chromophores located
dropped in the blue InGaN chip to yield a LED device for inside the host and the outside world is realized via stopcock
backlight display with high stability and wide color range. molecules of different shape, size, and nature (Figure 56A). The
Furthermore, green CsPbBr3 and red CsPb(Br/I)3 crystals were dynamics of the EnT processes in SBMs were presented in
used to prepare the composites that gave efficient and stable section 3.3, and some of the theoretical results were described
white LEDs.535 The polyhedral oligomeric silsesquioxane in section 2.2. It was demonstrated that it is possible to obtain
enhances the performance of a LED device based on well-organized dyes in zeolite L material in which FRET
CsPbBr3.536 The result was explained in terms of hole-blocking efficiency can be modulated by selecting suitable dye
properties of the silica-based layer, which keeps both electrons combinations (with a special attention to the spectral overlap
and holes located within the active perovskite layer for more of donor emission and acceptor absorption).16 Several levels of
efficient recombination. organization have been realized for zeolite L, extending from
A well-known problem of hybrid perovskites is their the interior of a given crystal to the channel entrances and the
external surface, as well as from the microscopic to the
instability when interating with water molecules (humidity in
macroscopic space domain.541 Recently, a variety of new dyes
air). The protection of MAPbBr3 nanocrystals from moisture
from the perylene diimide, terrylene, and quaterrylene family
was studied using silica coating.538 A novel method of
were tested for supramolecular organization in zeolite L
preparation of the SiO2 layer was proposed, since the
channels.385 Plugging the channel prevents the guests from
conventional silica formation method required water.539 On
escaping to the outside environment, which is essential for
the contrary, the new preparation method using toluene achieving long-term stability of such composites. The dye-
without ligand exchange yields remarkable photostable silica- zeolite L composites were also inserted into polymer matrices
encapsulated perovskite (94% photoluminescence quantum and maintained their optical transparency.
yield after 7 h of irradiation). MAPbBr 3 and MAPbI 3 In addition to zeolite L, another SBM for light harvesting and
perovskites were confined in a gyroidal mesoporous silica transport is the PMO, which is formed via polycondensation
template of various sizes from 5 to 12 nm.539 The crystal between organosilane precursors.542 The organization of the
structure and optical properties (absorption edge and emission chromophores in the PMO wall, along with appropriate guest
maximum) show a consistent change as the particle size is molecules in their meso-channels, make a versatile donor
reduced. Quasi-2D, 2D, and 3D perovskites (MAPbBr3) were scaffold for the EnT process (Figure 56B). It should be noted
stabilized in MCM41 matrix.540 All the formed composites that artificial antenna systems based on zeolites have also been
exhibit strong photoluminescence. The electroluminescence proposed for application in photovoltaics and other solar
signal was also observed and attributed to EnT from higher energy conversion devices. Other aspects of the potential use of
bandgap perovskites (quasi-2D, 2D) toward lower band gap SBMs in solar cells will be briefly presented in the section 5.3.
perovskites (3D). MAPbI3 perovskite nanocrystals (1−4 nm 5.2.1. Sensors. Host−guest systems based on SBMs can
diameter) were grown within periodically mesostructured films also be used as sensors. A sensor is defined as an entity
of TiO2 and SiO2.537 Tunable quantum confinement effects (molecule or material) that can detect changes in its
were observed due to the dimensions of nanocrystals environment, altering its properties as a result. Changes in
comparable to the exciton Bohr radius. For example, a large fluorescence intensity, in particular, are considered the most
blueshift of the emission intensity maximum was observed over effective tool for sensing applications due to their high
a wide range of 0.34 eV (Figure 55B). This result demonstrates sensitivity, easy visualization, and short detection response
BA DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

time. Sensors fabricated by encapsulating a fluorescent probe in


SBMs have demonstrated interesting properties such as high
thermal and mechanical stability, high brightness, photo-
stability, and relatively low cytotoxicity.8,543
With dependence on the analyte of interest, many types of
sensors can be found. pH Sensors are some of the most in
demand mainly due to their applications in biomedical
diagnosis.14,19,544−546 These sensors are fabricated with pH-
sensitive fluorophores encapsulated in or interacting with micro
or mesoporous materials. For example, spectroscopy techniques
have demonstrated the good pH sensitivity of a sensor based on
zeolite-β encapsulating 3-HF with a shell of amorphous silica
with fluorescein isothiocyanate (FITC, Scheme 8) mole-
cules.545 The experimental results demonstrated that this
sensor has high sensitivity, stability, and real-time response
for pH changes. While the caged 3HF did not show significant Figure 57. Emission spectra of Ru-MCM41 in water solutions at pHs
changes in the pH range of 5−8, the fluorescence signal of the from 1.07 to 12.92. Inset (up): emission intensity change with the pH
fluorescein embedded in the amorphous silica shell was value. Inset (down): fitted working plot (red line) in a pH range of
strongly affected, reaching its maximum value at pH = 8.545 4.22−6.64. Reprinted with permission from ref 549. Copyright 2014
Another study reported a sensor formed by mesoporous silica Elsevier B.V.
nanoparticles (MSNs) doped with R6G-lactam (Scheme 8) and
FITC.546 The R6G-lactam shows strong fluorescence in acidic radiative decay of Ru(II) and the quenching of its emission
media, while FITC exhibits the opposite behavior, exhibiting its signal. From pH 7.16 to 11.07, the emission signal is stable
maximum fluorescence intensity in basic media. This dual because the PIP ligands are present in their neutral forms and
colored sensor was characterized in the presence of sodium no effect on the fluorescence of the Ru(II) center was observed.
phosphate buffers of various pH (3.5−7.5). At acidic pH (3.5), However, when the pH is higher than 12, the deprotonation of
the emission spectrum has an intensity maximum at 552 nm the PIP ligands leads to an increase in their electron density,
(rhodamine-lactam intensity maximum), which decreases at further favoring the nonradiative decay pathways and thus
higher pH values with a simultaneous increase in the quenching the fluorescence of Ru(II).549 These results have
fluorescence signal at 515 nm (FITC intensity maximum).546 demonstrated that the most sensitive pH region for this sensor
This favorable inverse pH response improves the sensitivity of was 4.22−6.64, which could be related to the mesoporosity of
this sensor. The lysosome acidity range (3.5−5.5) overlaps with the MCM41 materials, as it favors the OH-diffusion to the
the optimal sensing range of this nanosensor (3.5−6.0), complex, resulting in quick quenching of its emission.549
suggesting that it could be used to monitor cell acidity and in Optical sensors for the detection of organic molecules and
lysosome-targeted cancer therapy.546 A similar system, formed metal cations have been recently investigated. For example, an
by hollow mesoporous silica nanoparticles (HMSNs) doped experimental study demonstrated the sensitivity of mesoporous
with rhodamine B isothiocyanate (RITC, Scheme 8) (pH organosilica films with 2D-hexagonal symmetry (p6mm)
insensitive) and FITC (pH sensitive), has been investigated to structure to the presence of the explosive TNT (Scheme
monitor intracellular pH.547,548 Its fluorescence spectrum shows 2).50 The system is formed by naphthalene-bridged silane
an increase in the intensity maximum of FITC (at 514 nm) at (BTPN, electron D, Scheme 8), which exhibits an ET process
high pH values, while the intensity maximum of RITC (at 576 with TNT (electron-A). The sensing rate was characterized by
nm) remains nearly constant.547,548 With the use of the change fluorescence experiments, as the adsorption of TNT molecules
of the ratio of the emission intensity of both dyes with the on these composites quenched the emission signal.50 The study
variation of pH values, a pH calibration curve was obtained showed that the sensitivity and selectivity of this sensor could
whose pH response ranges are 4.5−8.5. This pH sensitivity be improved by modifying the pore length and adjusting the
range is broader than those of systems based on MSNs (3.5− rate of TNT adsorption, respectively. Thus, the mesoporous
5.5) due to the larger pore size of HMSNs (2.9 nm vs 2.1 nm material directly affects the properties of the sensor. Similar
for MSNs) and the broader surface curvature distribution.547,548 behavior was observed for a system formed by mesoporous
An extension of the pH response between 3.2−9.0 was possible silica thin films containing a phenyl-substituted pyrene
by tuning the thickness and pore size distribution of these fluorophore (TKMPP, TKHPP, and TKSPP, Scheme 8) as a
systems, making them one of the best candidates for pH sensor of another nitroaromatic compound, 2,4-dinitrotoluene
sensors.548 Another report has shown a new pH sensing device (DNT, Scheme 8).550 The mesoporous architecture (wormlike,
formed by mesoporous MCM41 materials functionalized with a 2D- and 3D-hexagonal mesopores) in the presence or absence
novel Ru(II) complex of Ru(Bphen)2PIP (Bphen = 4,7- of the amphiphilic surfactant P123, and the mode of
diphenyl-1,10-phenanthroline and PIP = 2-phenyl-1H-imidazo- introduction of the fluorophore (bonded or impregnated on
[4,5-f ] [1,10]phenanthroline).549 When excited at 460 nm, the framework) influenced the detection range of the fabricated
these composites display an emission signal with an intensity sensor. In mesoporous films doped with pyrene dye and in the
maximum at 602 nm, which decreases with increasing pH presence of the surfactant P123, different efficiencies were
values (1.07−12.92) (Figure 57). observed according to the structure of the mesoporous films.
At low pH values (1.07−4.22), the PIP ligands of the Ru(II) Quenching levels of 45% and 9% were reported at 45 s for the
complex are protonated, inducing emission from the Ru(II) wormlike and 2D-hexagonal structures, respectively; while a
center. At even higher pH values (4.22−6.64), the PIP ligands lower quenching of 28% for the 3D-structures at 90 s was
undergo deprotonation, leading to an increase in the non- recorded.550 In fluorophore-bridged films and in the presence
BB DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

of surfactant, the reported quenching was lower for all sites was examined in the presence of different metal ions such
structures, reaching emission quenching efficiencies of 8.5% as Ag+, Na+, K+, Pb2+, Al3+, Mg2+, Zn2+, Cd2+, Hg2+, Co2+, Fe3+,
and 3% after 30 s for the wormlike and 2D-hexagonal Ni2+, and Cu2+, strongly detecting the last four cations, most
structures, with no detected changes for the 3D-hexagonal notably Fe3+, in the presence of which the emission signal was
structure. However, in the absence of P123, these composites strongly quenched.553 The selectivity of this system toward
show the best quenching performance, with 72% and 14% metal ions arises mainly from the exchange interaction between
quenching after 30 s for the wormlike and 2D-hexagonal these and zeolite A hosts. The fluorescence response range was
structures, respectively.550 The detection of other pollutants determined with different concentrations of Fe3+ (up to 1
such as ethylenediamine (En) is of great importance for mmol L−1), observing that the luminescence intensity of Eu3+
chemical industries due to its potential threat to human health. decreased with the increase in [Fe3+], reaching a total
Another report studied a selective sensor for En formed by quenching at 1 mmol L−1. The detection limit was 0.224
zeolite Y (ZY with terbium-acetylacetone (Tb(acacn)) mmol L−1.553 This value is higher than that reported for
complexes in their cavities.551 The Tb(acacn)/ZY composites another sensor composed of a hybrid mesoporous silica
showed considerable changes in luminescence color in the functionalized with a silane derivative porphyrin (TPP) and
presence of En vapor. Figure 58 shows the CIE chromaticity quinolone.554 In addition to detecting Fe3+, this hybrid material
is also sensitive to other trivalent metal ions such as Cr3+ and
Al3+. The presence of these ions simultaneously quenched the
TPP emission (650 nm) and enhanced the fluorescence of
quinolone (∼500 nm), causing a change of emission color from
red to green (Figure 59).554 The hybrid composites showed

Figure 59. Variation in the fluorescence intensity ratios (I650 nm/


I500 nm) of mesoporous silica (20 mg mL−1) with the concentrations
of Fe3+, Cr3+, and Al3+ ions in ethanol solutions. The inset shows the
fluorescence color changes. Reprinted with permission from ref 554.
Figure 58. CIE chromaticity coordinates of Tb(acacn)@ZY upon Copyright 2016 Wiley-VCH.
exposure to various solvent vapors for 10 min. Reprinted with
permission from ref 551. Copyright 2016 Elsevier B.V.
low detection limits of 0.08, 0.1, and 0.2 mmol L−1 for Fe3+,
coordinates in the presence of En (0.21, 0.22) and other Cr3+, and Al3+, respectively (Figure 59). The different response
solvent vapors (in the range from 0.26, 0.33 to 0.31, 0.49), degrees to Fe3+, Cr3+, and Al3+ allow the use of this material for
demonstrating the excellent selectivity of these composites for the simultaneous detection of these three ions.554
the detection of En.551 SBMs have also been used as humidity sensors. A recent
In addition to good selectivity, this sensor also shows high report demonstrated a remarkable dynamic change in the
sensitivity. To demonstrate it, experiments with different emission color of Ag-LTA zeolite in the presence of water.555
concentrations of En in the presence of other solvents such These composites were investigated following thermal treat-
as Et3N, n- and t-BuNH2, benzylamine, and methyaniline were ment (Figure 60). At room temperature, a broad band centered
performed, revealing a detection limit of 2% (volume ratio) at 600 nm was observed. However, after heating, the
Tb(acacn)ZY.551 A similar system formed by zeolite L fluorescence of the zeolite composites was different. At 473 K
functionalized with Tb-2-thenoyltrifluoroacetone (Tb-TTA) (approximately 20 water molecules per normalized unit cell),
complexes was used for luminescent sensing of dipicolinic acid the zeolite composites show the most remarkable changes with
(DPA, Scheme 8), which is the main constituent of many the presence of three well-defined emission bands at 560, 470,
pathogenic bacteria.552 The intensity of the Tb(III) lumines- and 390 nm. At 573 K (approximately 10 water molecules per
cence increased upon increasing the concentration of DPA due normalized unit cell), the emission band at 560 nm disappears
to the formation of a Tb-TTA-DPA complex. This sensor and the 470 and 390 nm bands blue-shifted. At 723 K
showed good selectivity and sensitivity, detecting DPA even at (approximately 1 water molecule per normalized unit cell), only
relatively low concentrations (0.05 μM).552 Another system an intense blue emission band is detected. Therefore, the fully
based on zeolite A with a lanthanide complex was used for the hydrated composites (20% water content) show a yellow
selective detection of metal ions.553 This composite was formed emission, whereas partially hydrated samples (17−3% water
by ytterbium-1,3-diphenylpropane-1,3-dione (Yb-DBM) com- content) present green and blue emission, and finally, the
plexes within the zeolite cavity and europium-1-(2-naphthoyl)- samples with less than 1% water content display only blue
3,3,3-trifluoroacetonate (Eu-NTA) ones bonded on the zeolite emission, thus demonstrating high sensitivity toward the
surface. The behavior of the [DBM-Yb-ZA]-NTA-Eu compo- presence of water molecules in the Ag-treated zeolite cages.555
BC DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 60. (A) Emission spectra of LTA(Li)−Ag1 after heat-treatment at different temperatures. (B) Color change of real samples with respect to
water content illustrated by the blue arrow. Reprinted with permission from ref 555. Copyright 2015 Royal Society of Chemistry.

Figure 61. (A) Schematic representation of the titanium clusters in the zeolite supercages. Reprinted with permission from ref 561. Copyright 2006
Wiley-VCH. (B) Illustration of the operation principles of DSSCs based on titanium-containing zeolites. Reprinted with permission from ref 562.
Copyright 2007 Wiley-VCH.

The systems with silica-coated chromophores were also tramethylrhodamine (CTMR, Scheme 8).557 A CTMR-SiO2
proposed in sensor applications, with the operation based on core was surrounded by concentric PTTA-Tb3+-SiO2. The long
EnT mechanism. Rhodamine-doped silica nanoparticles were emission lifetime of the donor (PTTA-Tb3+), 1.21 ms, was
linked to amphiphilic black hole quenchers to form a probe of reduced to 246 μs in the core−shell structure, which resulted in
reductive environments in living cells.556 Depending on the a very efficient EnT to CTMR with a quantum yield of 80%, in
number of quenchers adsorbed on each particle, the agreement with the results calculated from the areas of the
fluorescence lifetime decreased from 3.7 ns (no quenchers) corrected sensitized emission curves.557 DY-630-MI (Scheme
to 1.9 ns (16 quenchers per particle). The fluorescence 7) was covalently encapsulated in the core of ∼20 and ∼30 nm
quenching, due to FRET, did not occur in the presence of core−shell silica nanoparticles.558 An approximately 15-fold
reductive agents. As a result, these silica-based NP probes increase in the fluorescence intensity was reached, which is
showed more than a 10-fold increase in their fluorescence ascribed to restriction of the isomerization process in the dye’s
intensity in a reductive environment. Rhodamine dye was also
excited state and the protection from solvent interactions. FCS
doped within relatively large silica NPs (160 nm diameter).448
revealed that energy hopping between the dye molecules within
A very high fluorescence quantum yield and single lifetimes of
the single nanoparticles can occur. This process was found to
nanoseconds were reported despite the dye concentration
within each nanoparticle exceeding the typical value of be more efficient in smaller (∼11%) than in bigger (∼5%)
dimerization in water. The absence of concentration quenching nanoparticles.558 In another study, a sensitizer (octaethylpor-
effect was explained by the lack of water leakage through silica phyrin Pd complex) and an annihilator (9,10-diphenylanthra-
walls and the presence of CTAB surfactant inside the particles cene, Scheme 8) were coloaded into silica nanoparticles to give
as a splitting element among the dye molecules. This system, blue-emissive up-conversion nanoparticles.559 The sensitizer
which exhibits a super fluorescence, was suggested as a absorbs green light and transfers the energy to the triplet states
promising candidate for imaging and labeling applications. of the annihilators. When the triplet−triplet annihilation
Core−shell nanoarchitectures were used to study FRET process occurrs, it results in the final excitation of the blue-
between a luminescent Tb3+ chelate, N,N,N1,N1-[4′-phenyl- emitting singlet state. This silica-coated system showed low
2′2′:6,′2′-terpyridine-6,6′-diyl]bis(methylenenitrilo)tetrakis- cytotoxicity and was successfully used to label living cells with a
(acetate)-Tb3+ (PTTA-Tb3+), and an organic dye, 5-carboxyte- very high signal-to-noise ratio.
BD DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 62. (A) Schematic representation of ET from excited TPC1 to titania-containing mesoporous structures (not to scale). Reprinted from ref
564. Copyright 2011 American Chemical Society. (B) Current−voltage curves of DSSC with TPC1 dyes attached to titania nanoparticles and
TiMCM41 material (with 6.3% Ti doping) under 50 mW/cm2 illumination. Reprinted from ref 326. Copyright 2011 American Chemical Society.

5.3. Photovoltaics The subsequent contribution of the same group was devoted
Silica is an excellent insulator with a high dielectric strength to Ti-modified mesoporous molecular sieves.563 The material
(107 V/cm) and large band gap (8.9 eV). Therefore, its direct with isolated titanium atoms forming part of a framework of
use in photovoltaics as a photoactive material or charge mesoporous MCM41 silica (Ti/MCM41) has a high surface
transporter is rather limited. However, the very high specific area of approximately 880 m2 g−1. In the DSSC’s configuration
surface area of micro- and mesoporous SBMs, the doping with N3 sensitizer, this material results in a photocurrent
ability that can drastically modify their material properties, and density of 45 μA/cm2. Further studies using a similar concept
the possibility of using SBMs as various templates do provide were performed with the organic dye TPC1 (Scheme 4) as the
sensitizer.563 Two kinds of Ti-doped mesoporous structures
routes for the potential application of SBMs in photovoltaics.
were used: titanium atoms incorporated into the previously
Below, few examples of attempts to use silica materials in dye-
prepared MCM41 framework by the grafting method and small
sensitized solar cells (DSSCs), QDs and perovskite solar cells
titania nanoparticles used as one of the components in building
are presented.
the hexagonal mesoporous structure of SBA15.564 Figure 62A
A key element in DSSCs is the interaction of dyes with a
presents a schematic visualization of the two structures,
network of semiconducting metal-oxide nanoparticles (usually
showing titania domains approximately of 2 nm in size in the
TiO2).27,560 Such materials constitute a layer with a high
TiMCM41 grafted material. The dye content per Ti atom is
specific surface area, allowing sunlight to be efficiently
much better for the TiMCM41 material. Therefore, this hybrid
harvested by the attached dyes over distances of few microns. structure was tested in solar cells and compared with the
Moreover, TiO2 is also a medium that can transport electrons performance of DSSCs using standard P25 titania nano-
injected from the excited dye into its conduction band. particles.326 As seen in Figure 62B, the photocurrent density is
Therefore, interest in using SBMs in DSSCs has been low, approximately 60 μA/cm2 for 50 mW/cm2 illumination,
correlated with the successful doping of SiO2 frameworks which is the main reason that the total sunlight conversion
with TiO2. One of the first attempts was the incorporation of efficiency of the TiMCM41 cells is approximately 2 orders of
discrete clusters of TiO2 within the internal micropore space of magnitude lower than that of the cells with P25 particles. The
zeolite Y.561 The clusters contained only a few titanium atoms photocurrent per Ti atom is again comparable in both systems
(Figure 61A), with Ti loading of 4.8 wt %. In contrast to the (as in the previous studies with N3 dye), but the photocurrent
inactive zeolite Y matrix, all the zeolite encapsulated TiO2 per loaded TPC1 dye (and, thus, per absorbed photon) is much
species showed a photovoltaic response when irradiated with 1 lower in the MCM41-based devices. The ultrafast and fast
sun illumination (100 mW/cm2, spectrum of AM1.5 standard). dynamics of EI and dye regeneration are even slightly better in
The most efficient photovoltaic cell was based on a nitrogen- the TiMCM41 cells (some details are presented in the
doped system (to enhance visible-light absorption) and gave a discussion of electron transfer, section 3.2.1.2). Therefore, the
photocurrent of 5.8 μA/cm2 when used with an iodide main reason for the overall low photocurrent in TiMCM41 was
electrolyte (which is commonly used in DSSCs).123 Later on proposed to be due to slow ET to the anode. Since electrons
the popular ruthenium sensitizer cis-bis(isothiocyanato) bis- need to hop between titania domains in the insulating SiO2
(2,2′-bipyridyl-4,4′-dicarboxylato ruthenium(II) (N3, Scheme material, the dynamics of this process might be insufficient to
8) was used for several titanium-containing zeolites and compete with electron recombination to the electrolyte
microporous molecular sieves.562 The best results were (occurring on the time scale of milliseconds to seconds),
obtained for nanocrystalline Ti/Beta zeolite, which gives a even at zero bias voltage (short-circuit conditions).
photocurrent of approximately 100 μA/cm2. Figure 61B shows Ti-doped zeolites or MCM41 materials were also tested with
the working principle of such a solar cell, emphasizing the need dyes that do not form covalent bonds with Ti atoms. The
for electron hopping between Ti traps as a possible mechanism emission of 7-HQ (Scheme 3) was observed to be partially
for electron transport to the anode. In the same solar cell quenched in TiMCM41 material with 3% or 6% titania content
configuration, P25 TiO2 nanoparticles (standard in DSSC and completely quenched in a pure titania support.200 The
architectures) give a much higher short-circuit current density ultrafast dynamics in these systems are presented in section
of 4.2 mA/cm2. However, the photocurrent density corrected 3.2.2.2. t-St (Scheme 4) was incorporated into MOR zeolite on
by the Ti content was calculated to be similar to the value for which TiO2 nanoclusters were deposited, and it was observed
Ti/Beta zeolite. that the presence of titania stabilized the dye radical cation.358
BE DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

This effect is attributed to the capture of the ejected electron by NaOH, giving a photocurrent density of 0.3 mA/cm2.
the semiconductor domains. Electrolyte-mediated charge transport between the intercon-
As mentioned above, the all-silica parts with high bandgap nected QDs in zeolite was proposed, with electron transfer
prevent fast electron transport in DSSCs. From the perspective from S2− anions in the solution to the hole in the valence band
of electron migration, materials made exclusively from TiO2 but of the CdS QDs. Interestingly, in the analogous configuration
maintaining the mesoporous structure of a silica template could (in the photoanode), PbS did not show any photovoltaic
be much more efficient. The potential advantage of this concept response. However, it was successfully used in the counter
is that a large crystalline titania network might transport electrode. In this configuration with both CdS-loaded zeolite on
electrons even better than a layer of sintered titania the photoanode and PbS loaded on the photocathode, the
nanoparticles, having many trap states at grain boundaries. overall photocurrent increased to almost 1 mA/cm2. The
An alternative nanoporous TiO2 material was prepared using a proposed energetic scheme and electron transfers for this
silicate SBA15 framework as a template.565 After the growth of arrangement are presented in Figure 64.
the TiO2 nanostructure, the silicate component was removed
from a NaOH solution. The photocurrent of a DSSC with this
material, sensitized with ruthenium dye ditetrabutylammonium-
cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato)
ruthenium(II) (N719, Scheme 8), is as high as 16.7 mA/cm2.
An analogous solar cell made with standard P25 titania particles
as a control exhibits approximately two times lower efficiency.
More recently, mesoporous TiO2 single crystals were prepared
from a framework of silica spheres, and all-solid DSSCs were
constructed entirely below 150 °C.566 The silica template was
finally removed by selective etching in aqueous NaOH to
recover the mesoporous TiO2 crystal product. This meso-
porous TiO2 gives a photocurrent higher than 6 mA/cm2 when
sensitized with the organic dye D102 (Scheme 8).
An interesting new concept for photovoltaics using zeolites,
called FRET-sensitized solar cells, was also proposed and is
illustrated in Figure 63.393 Here, light is absorbed over a broad Figure 64. Diagram of the energy levels in the most efficient
configuration of quantum dots solar cell in zeolite Y. Reprinted from
ref 568. Copyright 2011 American Chemical Society.

Finally, the optoelectronic properties of metal halide


perovskite material of huge interest in photovoltaics570 have
been also recently tested in interaction with SBMs for
perovskite-based solar cells improvement. Methyl ammonium
(MA) lead halide perovskites were recently incorporated inside
mesoporous silica templates.571 The average fluorescence
lifetime decreases from hundreds of nanoseconds in the bulk
to tens of nanoseconds in the silica template. The lifetime also
Figure 63. Schematic illustration of a FRET-based DSSC. The decreases when the pore size was decreased from 7.1 to 3.3 nm,
composite consists of nanochannels containing two types of accompanied by a blue shift of the perovskite emission band.
chromophores (blue and green) and stopcocks (red). Reprinted Interestingly, it was found that the confinement in the rigid
from ref 393. Copyright 2012 American Chemical Society. silica template protected the perovskite from various types of
degradation that occur in the bulk.570 Although such silica-
spectral range in the dye-zeolite antenna composite. The coated perovskites are not applicable in photovoltaics, they
excitation energy migrates by radiationless transfer (FRET) have been proposed as useful reference materials for
along the inserted dye in one direction (because of the specific comparison with other perovskites or as functional materials
ordering of the dyes) and reaches the active medium to drive in all-solid-state light-emitting diodes (see previous section on
the charge-separation process. The illustration in Figure 63 photonics applications).
shows that the excited stopcock molecules (at the end of SBMs have been recently proposed as alternative meso-
antenna system) transfer their energy via FRET to the porous scaffolds in perovskite solar cells. So far, the most
semiconductor (silicon). Another version of FRET-sensitized typical material used in this context in perovskite solar cells was
solar cells was proposed in conjunction with organic solar cells, TiO2, which assists in electron transport (like in DSSCs).
where the active medium is polymer.567 In such solar cells, the However, it was shown that nanostructured materials with large
common drawback of the active material is the limited spectral bandgap (like Al2O3) can also be used in efficient perovskite
range of high absorption of light. Zeolite-based antenna systems solar cells because the electrons are easily transported to the
can enhance light harvesting ability for photons of higher electrode within perovskite itself, without mediation of the
energy than that of the active medium. scaffolds.572 Therefore, SiO2 nanoparticles were first tested.573
CdS and PbS QDs were encapsulated in Y zeolites and tested Particles of different diameters (from 15 to 100 nm) were
in solar cell configurations.568,569 CdS-loaded zeolite Y films probed, and the best results were obtained for 50 nm silica
grown on indium tin oxide (ITO) glass show a photovoltaic NPs: sunlight conversion efficiency 11.45% with respect to
response in an electrolyte solution consisting of Na2S and 10.29% for the device with TiO2 scaffold layer. The lower
BF DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 65. (A) Illustration of perovskite penetration into the scaffold silica layer and its dependence on the SiO2 nanoparticles sizes. Reprinted from
ref 573. Copyright 2014 American Chemical Society. (B) Luminescence decays of MAPbI3 films of different scaffold materials. Reprinted with
permission from ref 576. Copyright 2016 Elsevier B.V.

performance of smaller silica NPs was explained in terms of not should be multifunctional and should possess the ability to on
sufficient infiltration of perovskite into the scaffold layer (Figure demand switch on and off certain functions.587−589 Due to their
65A). SiO2/TiO2 hollow nanoparticles improve the perform- versatile properties for drug delivery, including biocompati-
ance with respect to pure TiO2 ones due to higher bility, high pore volume, tunable pore size, and versatile
photovoltage.574 The highest efficiency is 14.7% for MAPbI3 chemistry for surface functionalization, mesoporous silica
perovskite and Al doping in the ET layer. Silica nanorods with nanoparticles (MSNs) are an attractive system for the
controlled aspect ratios were then used as scaffold layers.575 development of drug delivery nanocarriers. The high pore
The highest conversion efficiency (13.9%) was obtained for volume and the ability to tune the pore size allow high drug
relatively long SiO2 nanorods with an aspect ratio of 5. In loading of both small-molecule drugs and large biomacromo-
comparison to nanospheres of 80 nm diameter, nanorods lecules.25,590 Furthermore, the surfaces of these nanoparticles
provide better perovskite infiltration. Even higher efficiency can be easily functionalized to allow the pores to be gated by
(16.2%) was achieved for commercial SiO2 nanoparticles molecules that are responsive to external stimuli (e.g., light,
(Degussa, A200).576 This system is better than those with heat, pH, redox, and magnetic fields, to name just a few
mesoporous TiO2 or Al2O3 scaffolds due to better photovoltage examples), most notably pH changes and light irradia-
and fill factor of the cells. Moreover, hysteresis for SiO2-based tion.25,590−592 To develop such systems, a great detailed
perovskite solar cell is also less severe, and the luminescence knowledge of their behavior upon cell internalization and
decay is the longest for silica layers (Figure 65B). during the triggered release is needed.
SBMs were also tested in planar perovskite cells. Using silica- A number of mesoporous silica drug delivery systems has
coated gold nanorods improves the efficiency from 15.6% to been developed to realize controllable drug release by
17.6%.577 The effect is attributed to localized surface plasmon employing different kinds of pH- and redox-sensitive pore
resonance of gold (section 3.3.4). Not only increased incident capping architectures that can be switched on to release drug
light trapping (higher photocurrent) but also improved molecules from the mesoporous structure.25,590,592−596 In many
transport and collection of charge carriers (better fill factor) cases, the extracellular environment around solid tumors
is observed in the devices with silica-coated gold nanorods. becomes acidic as a result of the Warburg effect, associated
Similar efficiency enhancement (from 15.85% to 17.6%) was with the increased production of lactate by the cancer cells, thus
reported upon adding hexagonal mesoporous silica islands to allowing for selective and targeted drug release.597 A recent
planar junction perovskite solar cells.578 Instead of forming a example is the introduction of a targeting polymer, poly-
continuous scaffold layer, wormholelike hexagonal mesoporous (ethylene glycol)−folic acid (PEG−FA), on the surface of
silica form random islands on top of a compact TiO2 blocking polydopamine (PDA)-modified MSNs.598 These nanoparticles
layer. This improvement is due to increased light trapping effect were employed as a drug delivery system loaded with
and pinhole-blocking effect. doxorubicin (DOX, Scheme 8). In this system, the pH-sensitive
PDA coating serves as a gatekeeper. In vitro release
5.4. Drug Delivery
experiments have shown pH-dependent, sustained drugs release
Drug delivery is a process in which several fundamental fields of profiles that could enhance the therapeutic anticancer effect and
modern science converge.579 With the development of new and minimize the potential damage to normal cells. Furthermore,
advanced nanocarriers, it is at the frontier of nanotechnology the nanoparticles achieved significantly high targeting efficiency,
and material chemistry and its underlying goal of bringing as demonstrated by the in vitro cellular uptake and cellular
therapeutic or imaging agents to the desired site is related to targeting assays.598 An MSN functionalized with disulfide snap-
biology and medicine.579,580 The delivery process can be tops has been developed showing high drug loading and
considered a succession of several steps: loading, delivery, selective intracellular drug release in response to the redox
targeting, and release. The medical applications of nano- potential.599 These nanoparticles, when loaded with a Hoechst
technologies have propelled the development of various types fluorescent dye, release their cargo exclusively intracellularly
of drug-loaded nanocarriers, such as liposomes, cyclodextrins, and stain the nuclei of macrophages. The particles loaded with
micelles, polymers, and metal−organic framework moxifloxacin kill F. tularensis in macrophages in a dose-
(MOFs).579,581−586 In the ideal case, nanoparticle-based dependent manner. In a different strategy, silylated precursors
delivery systems are expected to accumulate in the required with molecular recognition sites based on triazine and uracil
organ or tissue and simultaneously penetrate target cells to fragments were grafted onto MSN surfaces.600 After loading
deliver the bioactive agent. Any smart drug delivery system with cargo molecules (dyes or drugs), the nanoparticles were
BG DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Figure 66. (A) Schematic presentation of the synthesis concept and drug release of Pt@PAA-MSNDOX. In vitro (B) Pt(II) and (C) DOX release
behaviors of Pt@PAA-MSNDOX in aqueous solutions with and without 0.9% NaCl at pH 7.4 and 5.5, respectively. Reprinted with permission from
ref 601. Copyright 2016 Royal Society of Chemistry.

successfully capped by H-bonding to the complementary caps, anchoring folic acid by the CuAAC click reaction, whereas
which were prepared by covalently coupling a uracil or adenine controlled delivery was performed by clicked azobenzene
unit to the bulky β-CD. The pH sensitivity of these systems was fragments.607 The azobenzene groups can obstruct the pores of
demonstrated by the release of dye molecules (propidium the nanoparticles in the dark, whereas upon irradiation by the
iodide or RhB) at acidic pH. The particles were stable at neutral UV or blue light, their trans-to-cis photoisomerization induces
pH, showing no premature release and a controlled release disorder in the pores, enabling the delivery of the cargo
triggered by an acid stimulus. Cytotoxicity experiments on molecules. The on-command delivery was proven in solution
human breast cancer cells using nanoparticles loaded with by dye release experiments and in vitro by DOX delivery.607
camptothecin (CPT, Scheme 8) showed a decrease in cell The added value of the folic acid ligand was clearly evidenced
viability similar to that induced by the drug alone, confirming by the difference in cell killing induced by doxorubicin-loaded
the release of the active compound after pore opening in the nanoparticles under blue irradiation depending on whether the
acidic lysosomes. A much smaller amount of the chemo- particles featured the clicked folic acid ligand or not. However,
therapeutic agent CPT (40 times less) than that of the free UV light activation is not easily applicable for biology, since
drug was loaded into these nanoparticles.600 A dual drug one-photon irradiation has limited tissue penetration, may
codelivery system based on poly(acrylic acid)-modified MSNs induce phototoxicity, and lacks spatial accuracy. To overcome
for the combination of DOX and cisplatin was also this limitation, recent studies have reported on the use of NIR
demonstrated (Figure 66).601 The cisplatin was conjugated two-photon-triggered drug release.608−610 To date, a variety of
with carboxyl groups to form pH-responsive cross-linking shells NIR light-responsive drug delivery systems based on different
after the encapsulation of DOX in the mesopores of the silica mechanisms have been developed for controllable drug
particle. As a result of the pH-sensitive drug release, 70% of the delivery.610 The photothermal effect, which normally utilizes
cisplatin and 80% of the DOX are released at pH 5.5, while only red or NIR-absorbing agents to generate heat from NIR optical
16% of the cisplatin and 25% of the DOX are released at pH energy, may be employed to trigger drug release from
7.4.601 photothermal-responsive drug delivery systems.610−612 Upcon-
In a different approach, a novel enzyme-based cap system version nanoparticles, which can be excited by multiple NIR
that is directly combined with a targeting ligand via bio- photons and emit a single high-energy photon with a shorter
orthogonal click chemistry was developed for MSNs.602 This wavelength, even in the UV region, have also been used for
capping system is based on the pH-responsive binding of an NIR light-responsive drug release.613 MSNs functionalized with
arylsulfonamide-functionalized MSN and the enzyme carbonic azobenzene and containing a two-photon fluorophore demon-
anhydrase (CA). An unnatural amino acid (UAA) containing a strated the triggered release of cargo molecules.614−616 The
norbornene moiety was genetically incorporated into CA, release was achieved by NIR light irradiation because the two-
allowing the site-specific bio-orthogonal attachment of very photon fluorophore emit 420 nm light, which matches the
sensitive targeting ligands such as folic acid and anandamide. azobenzene absorption band, leading to successful isomer-
This specific attachment leads to specific receptor-mediated cell ization of the azobenzene gates.616 Two-photon-sensitive
and stem cell uptake. The study reported the successful delivery mesoporous organosilica nanocarriers (M2PS) were synthe-
and release of the chemotherapeutic agent actinomycin D to sized via the co-condensation of a silica precursor and a two-
Kleihauer-Betke (KB) cells.602 photon electron donor (2PS).612 First, one- and two-photon-
MSNs can also be functionalized so that their pores are gated actuated cargo (DOX) release via the photoreductive cleavage
by light-responsive molecules.603,604 Several strategies have of the disulfide nanogates of the MSN using M2PS-35 and
been proposed and realized for light-activated cargo release M2PS-16 was successfully monitored in aqueous solutions.
from these materials.603,605,606 The initial strategy was based on Furthermore, the cellular uptake in Michigan Cancer
the use of a UV−visible light. Bis(clickable) mesoporous silica Foundation 7 (MCF-7) cells was demonstrated via two-photon
nanospheres (ca. 100 nm) were obtained by the co- fluorescence imaging of the nanoparticles, which were then
condensation of TEOS with variable amounts of two clickable applied successfully for drug delivery in cells (Figure 67).612
organosilanes (2−5% each) in the presence of CTAB.607 These Spatially confined cyanine-anchored silica nanochannels
nanoparticles could be functionalized easily with two loaded with chemotherapeutic DOX were synthesized for
independent functions using copper-catalyzed alkyne−azide light-driven synergistic cancer therapy.617 The cyanine dye is
cycloaddition (CuAAC) to transform them into nanocarriers present within the nanochannels in a J-type aggregation
bearing cancer cell targeting ligands with the ability to deliver conformation, and DOX is encapsulated via π−π interactions
drugs on demand. Active targeting was made possible by with the cyanine dye. Under NIR light irradiation, the loaded
BH DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

from the nanorod surface and thus triggering the release of the
loaded DOX or G3-Pt. As evidenced by both in vitro and in
vivo experiments, CMSNR-B-PEG with either DOX or G3-Pt
loading offers remarkable synergistic therapeutic effects in
cancer treatment owing to the on-demand release of
therapeutics specifically in the tumor under light irradiation.619
Ce6-gold composites in mesoporous silica shell for drug release
were also studied,417 and they were shortly presented in section
3.3.4.
The combination of pH and light as external stimuli has also
been reported. MSNs have been synthesized and loaded with
both aluminum chloride phthalocyanine (AlClPc) and cisplatin
as combinatorial therapeutics for treating cancer.620 The
intracellular uptake and cytotoxicity were evaluated in human
cervical cancer (HeLa) cells by confocal laser scanning
microscopy (CLSM) and 3-(4,5-dimethylthiazol-2-yl)-5-(3-
carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium
(MTS, Scheme 8) assays, respectively, showing that the
mesoporous materials can be readily internalized by HeLa
cells.620 The cytotoxicity experiments demonstrated that after
light exposure, the combination of AlClPc and cisplatin
compounds in the same platform potentiates the toxic effect
Figure 67. Images of (A) M2PS-35 and (B) M2PS-16 using two- against HeLa cells in comparison to the effects of the control
photon fluorescence showing the cellular uptake of the nanoparticles AlClPc-MSN and cisplatin-MSN materials.620 A similar
(Scale bar 10 mm). (C) Two-photon triggered DOX delivery using approach has been applied to the synthesis of a dual pH-
M2PS-35 and M2PS-16 silica-based gates. Reprinted from ref 612. responsive MSN-based drug delivery system for synergistic
Copyright 2015 American Chemical Society.
chemo-photodynamic therapy.621 By grafting histidine onto the
silica surface, acid-sensitive PEGylated tetraphenylporphyrin
particles exhibit enhanced photothermal conversion efficiency zinc (Zn-Por-CA-PEG) was used as a gatekeeper to block the
due to the maximized nonradiative transition of the J-type nanopores of the particles via the metallo-supramolecular-
aggregates. These trigger light-driven drug release through the coordinated interaction between Zn-Por and histidine. This
destabilization of temperature-sensitive π−π interaction and gatekeeper is stable enough to prevent the loaded drug from
cause the effective intracellular translocation of DOX from the leaching out in healthy tissue. However, at the extracellular pH
lysosomes to the cytoplasm via reactive oxygen species- of cancer, the conjugated acid-sensitive cis-aconitic anhydride
mediated lysosomal disruption, thereby causing potent in vivo between Zn-Por and PEG is cleaved, and the surface of Zn-Por
hyperthermia and intracellular trafficking of the drug into the becomes positively charged to facilitate cell internalization.621
cytoplasm at tumor sites.617 A red light (660 nm)-responsive Due to the removal of the gatekeeper, the metallo-supra-
drug delivery system based on cyclodextrin (CD)-gated MSNs molecular-coordination disassembles in intracellular acidic
containing a photodynamic therapy (PDT) photosensitizer Ce6 microenvironments to release the carried drug and Zn-Por.
(Scheme 5) was reported.618 In water, Ce6 can be excited by The photosensitivity of Zn-Por further enables the combination
red light to generate singlet oxygen, which can further cleave of chemotherapy and photodynamic therapy.621 The idea of
the sensitive linker to trigger the departure of CD and the dual-stimulus synergy has been further expanded to a new
release of cargo. In vitro release experiments have shown the nanocomposite using reduced graphene oxide (rGO) and
cargo to be released from MSNs by irradiation with red light mesoporous silica−DOX/hydroxyapatite (HA).622 Owing to
and then to spread into the entire cell. The relatively low power the relatively high loading capacity of the anticancer drug DOX,
density (0.5 W cm−2) of the excitation light together with the the biomimetic capping of nontoxic and pH-dependent
short irradiation time (1−3 min) result in a low light dose (30− biodegradable HA on the surface of the composite and the
90 J cm−2) for the drug delivery.618 A novel light-responsive efficient photothermal conversion effect of rGO under NIR
drug delivery platform based on Ce6-doped mesoporous silica irradiation, the nanocomposite provides a pH−light dual-
nanorods (CMSNRs) was developed for on-demand light- triggered drug release. The in vitro experimental results have
triggered drug release.619 In one such architecture, the indicated that the chemotherapeutic effect of DOX and the
nanorods are coated with bovine serum albumin (BSA) protein hyperthermia induced by rGO under NIR light irradiation kill
via a singlet oxygen sensitive bis(alkylthio)alkene (BATA, HeLa cells effectively.622
Scheme 8) linker and then modified with polyethylene glycol Finally, mesoporous SBMs can be used both as drug
(PEG).619 The obtained CMSNR-BATA-BSA-PEG, termed nanocarriers and as “reactors” for the noncovalent assembly
CMSNR-B-PEG, acts as a drug delivery carrier loaded with of synergetic entities (Figure 68). The first such example of
either small drug molecules such as DOX or larger macro- noncovalent assembly between FA and a porphyrin [tetrakis(4-
molecules such as cisplatin predrug-conjugated third-generation carboxyphenyl)porphyrin, TCPP] in mesoporous silica par-
dendrimer (G3-Pt), both of which are sealed inside the ticles has been reported recently.623 The assembly process does
mesoporous structure of the nanorods by BSA coating. Upon not occur in aqueous solutions but occurs exclusively in the
660 nm light irradiation with a relatively low power density, interior of the mesoporous silica material that entraps both
CMSNRs with intrinsic Ce6 doping generate singlet oxygen to components. The assembly shows enhanced and preserved
cleave the BATA linker, inducing the detachment of BSA-PEG fluorescence in the FA and TCPP spectral regions, respectively,
BI DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(monomers or aggregates) is demonstrated in hundreds of


reports using time- and space-resolved laser-based techniques.
Not of less importance, theoretical approaches have been
applied to explore solvation and guest interaction with the
SBMs. Due to the related composites complexity and large
number of atoms (host + guest + solvent) to consider, the
available theoretical tools are not as sophisticated. However,
relevant information on the composites and nanosolvation
could be extracted, suggesting new experiments. We have
further reviewed the development of different applications
associated mainly with the use of light to trigger the desired
process in these composites. These applications include
Figure 68. Schematic representation showing the concept of PDT photocatalysis, nanosensors, photonics, photovoltaics, and
using FA/TCPP assembly released from mesoporous silica material to drug delivery systems, demonstrating the versatility and
kill cancer cells. Reprinted with permission from ref 623. Copyright significance of these materials for modern science and
2017 Wiley-VCH. technologies. Clearly, thanks to the advancements in ultrafast
and single-molecule spectroscopic techniques, computational
along with the ability to photosensitize the formation of singlet methods, and theoretical models, the understanding of many
oxygen slightly better than the free photosystem can. The photophysical processes in SBMs, such as solvation dynamics,
assembly has demonstrated high potential for the photo- proton, electron, and energy transfer has increased in the past
inactivation of KB cancer cells over expressing the folate decade. The reviewed works have provided a wealth of
receptors, compared with the potential of free TCPP, most paramount information on the excited-state behavior of organic
likely because of the FA targeting effect.623 molecules and metal complexes interacting with SBMs, in
In this part, we highlighted the versatility of SBMs for the particular, depicting photoinduced dynamic processes as
development of various and diverse applications. The specific mechanistically different from those in homogeneous solutions.
features of SBMs, such as high surface area, high thermal and Yet, a certain discrepancy exists between the theoretical and
mechanical stability, good and selective adsorption capacity, experimental works. On one hand, the details of the predictions
and the presence of active sites make these materials one of the of models for solvation or vibrational correlation require
most used in the field of catalysis (section 5.1). The design of experimental verifications. On the other hand, the complex
new sensors (section 5.2) and their encapsulation leading to the systems investigated experimentally for the possible applica-
formation of composites of high photostability and brightness is tions are usually still too complicated for the current calculation
another field under development. The use of these systems resources and methods. Developing more precise theoretical
allows development of smart sensor devices. Light collection treatments that describe complex kinetic behavior of excited
and transmission, which can be realized using 1D oriented species within SBMs are very important to get a picture at
hosts (most often zeolite L) with properly designed atomic and orbital levels. Nevertheless, the combination of the
chromophores inside, acting as antenna-like systems, is another theoretical and experimental approaches has helped to clarify
field of nanophotonics (section 5.2). Increased interest has the SBMs considerably from the photodynamical point of view.
been recently shown to new sources of light based on SBMs. In The ultrafast spectroscopic techniques still have experimental
dye-sensitized solar cells, micro- and mesoporous silica limitations and technical issues, related to the nature of the
materials have been used instead of semiconducting titania samples, which are in suspensions or solids, and are highly
nanoparticles (Section 5.3). Silica-coated gold nanostructures heterogeneous, to be overcome for providing high quality
have been tested to enhance the light absorption by the active experimental data. Technological advances should allow for the
layer of solar cells. Mesoporous silica scaffolds in perovskite development of more sensitive techniques with better time and
solar cells are perhaps the most successful use of these materials spectral and spatial resolution that can monitor weakly bonded
in photovoltaics so far, playing a totally passive role (electrons guest-SBMs. This in turn will open the possibility to better
are transported via perovskite material, unlike the typical characterize the molecular interaction in SBM complexes with
architectures with titania layers). Finally, mesoporous SBMs application in photonics, catalysis, sensors, or drug delivery.
represent an attractive alternative to the conventional drug From the applications point of view, the meso- and
delivery systems (section 5.4). The controlled release offered microporous SBMs have a unique set of properties, such as
by these materials has many advantages such as high release high surface area, high thermal and mechanical stability, good
efficiency, precise control of the dosage form for prolonged and selective adsorption capacity, presence of active sites, and
periods, and decreased toxicity. Surface modifications open new ability for surface modifications. These features are paramount
possibilities for controlled drug release and targeting drug for the development of a variety of new smart devices with
delivery. In this section, two of the more investigated applications in a broad range of technologies. With the
approaches (light and pH) for external stimuli-responsive improvement of our fundamental knowledge of photoinduced
drug delivery have been reviewed. processes within the related composites, we should be able to
address further challenges in related fields, such as selective
6. GENERAL CONCLUSION AND OUTLOOK control and better yield of photoinduced reaction (photo-
In this review, we have commented on and discussed the catalysis), better sensitivity and selectivity (sensors), more
confinement effect of SBMs on the spectroscopy and dynamics efficient light absorbing and converting devices (photonics), or
of guests, ranging from classical organic dyes to drug molecules improved targeting and selectivity to enhance cellular/tissue
and quantum dots. The basic concepts of confining molecules recognition to specifically match only the receptors present in
to change their photobehavior and related molecular state diseased cells (drug delivery). Overcoming these challenges will
BJ DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

open new avenues in the SBMs science and technology. For ultrashort pulse propagation, and supercontinuum generation. After
example, limited brightness and photostability of many probes completing his Ph.D. degree (in 2003 at AMU), he started to study
are one of the major constraints in using them in sensing in excited state proton transfer process and photochromism phenomena.
chemical and biological systems. Overcoming this limitation by From 2009 to 2011, he was a postdoc in the group of Prof.
trapping into or bonding them to SBMs would allow for Abderrazzak Douhal at Universidad de Castilla La Mancha in Toledo,
improved spatial resolution in confinement and tracking studies Spain, under the Marie Curie Fellowship. He investigated photo-
getting higher performence imaging for better understanding chromic molecules and dyes for solar cells encapsulated in zeolites and
and smarter design. Such composites with well-defined sizes, mesoporous silica sieves. In 2013, he completed his habilitation degree
surface hydrophobocity/hydrophilicity, sensitivity to changes in at AMU. His current scientific interest are dye-sensitized and
the local environment, and having strongly polarized emission perovskite solar cells, especially studied by time-resolved laser
would also be useful for probing the accessible internal spectroscopy techniques. (ResearcherID: J-6666-2012; ORCID:
dimensions of confining nanostructures, including canals of orcid.org/0000-0003-1882-6022).
biological moecules. Advanced ultrafast vibrational, electronic Abderrazzak Douhal (https://www.uclm.es/profesorado/adouhal/
and X-ray spectroscopy, and 4D electron microscopy are some douhal.htm), graduated (1982) from the Faculty of Sciences at Kadi
of the modern tools to explore the intimate interactions in the Ayyad University (KAU, Marrakech, Morocco) and is a Professor of
formed composites at atomic and short-time scales.624−635 With physical chemistry at the University of Castilla La Mancha (1993,
these powerful techniques, one can explore for example the UCLM, Toledo, Spain). He has received his Ph.D. degree in chemistry
time evolution of the electric field effect generated by the from KAU after a research period at the Institute of Physical
doping metal in the framework on the composite robustness Chemistry “ROCASOLANO” at “Consejo Superior de Investigaciones
and performance. While with the current knowledge on ́
Cientifica, CSIC”, Madrid. During 1990−1992, he was a postdoctoral
photophysics and photochemistry one can predict the photo- student (MONBUSHO and JSPS fellow, Japan) at the Institute for
behavior of a given molecule in solution when the properties of Molecular Science (Okazaki, Japan). In 1993, he worked as a research-
the solvent are known, we hope that, in the near future, similar associate at “Laboratoire de Photophysique Moléculaire, LPPM”
understanding related to SBMs will come, knowing the basic (University of Paris-Sud/CNRS, France). He was a visiting researcher
properties of silica material (morphology, size, doping, etc.) will at California Institute of Technology (Caltech) for several periods
enable one to foresee the photoresponse of a given from 1995−2000, collaborating with Prof. Ahmed H. Zewail. Since
encapsulated organic or inorganic guest in the composite. 1998, he is heading the Femtoscience and Microscopy research group
at the UCLM, focusing his research on the study of photoevents in
AUTHOR INFORMATION condensed phase and advanced materials (silica-based materials,
Corresponding Author MOFs, COFs, HOFs, and perovskites). ResearcherID: L-4940-2014;
*E-mail: abderrazzak.douhal@uclm.es. Tel: +34 925 265717. ORCID: 0000-0003-2247-7566.
ORCID
Marcin Ziółek: 0000-0003-1882-6022 ACKNOWLEDGMENTS
Abderrazzak Douhal: 0000-0003-2247-7566 This work was supported by the MINECO (MAT2014-57646-
Notes P, Consolider Ingenio 2010: CSD2009-0050 MULTICAT, and
CEI CYTEMA: CEI15-08-5), JCCM (PEII-2014-003-P), and
The authors declare no competing financial interest. NCN (National Science Centre, Poland, 2015/18/E/ST4/
Biographies 00196). N.A. thanks the CYTEMA program for the
postdoctoral grant. We thank Dr. M. R. Di Nunzio for her
Noemi ́ Alarcos studied chemistry at the University of Castilla-La help in drawing the chemical molecular structures.
Mancha (UCLM, Spain) and obtained her diploma in 2010. She
obtained (2015) her Master’s and Ph.D. diplomas in (Physical)
DEDICATION
Chemistry from the UCLM under the supervision of Prof.
Abderrazzak Douhal. Since 2015, she is continuing her research as a This work is dedicated to the Memory of Ahmed H. Zewail.
postdoctoral student in the same group thanks to a CYTEMA grant.
Her main research interest is focused on the fast and ultrafast ABBREVIATIONS
dynamics of photoinduced processes of dyes and drugs in solutions General Abbreviations
and within silica-based materials. ORCID: orcid.org/0000-0003-1350-
9036. SBMs silica-based materials
ICT intramolecular charge transfer
Boiko Cohen studied chemistry at the Sofia University, Bulgaria, and TICT twisted-intramolecular charge transfer
obtained his diploma in 1997. He obtained his Ph.D. (2003) from the FRET Förster (fluorescence) resonance energy transfer
Tel Aviv University. After postdoctoral research at the Ohio State 2D-IR two-dimensional infrared spectroscopy
University and Northwestern University, he joined University of DFT density functional theory
Castilla−La Mancha in 2007 as Ramon y Cajal Fellow. He has been an TD-DFT time-dependent density functional theory
Associate Professor at the University of Castilla−La Mancha since MOs molecular orbitals
2012. His main research interest is the fast and ultrafast dynamics of HOMO highest occupied molecular orbital
photoinduced processes in solution and in confining chemical and LUMO lowest unoccupied molecular orbital
biological media. (ResearcherID: C-4381-2008; ORCID: orcid.org/ HOC hydrophobic cavities
0000-0002-5400-4678). HIC hydrophilic cavities
Marcin Ziółek is currently an Associate Professor at the Faculty of RHOC rugged pores of the hydrophobic cavities
Physics, Adam Mickiewicz University (AMU) in Poznań, Poland. His PT proton transfer
first scientific topics were femtosecond pump−probe spectroscopy, ESPT excited-state proton transfer
BK DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

GSPT ground-state proton transfer k1, k2 (k‑1, k‑2) forward (reverse) proton-transfer rate con-
ESIPT excited-state intramolecular proton transfer stants
GSIPT ground-state intramolecular proton transfer τ fluorescence lifetime
UV ultraviolet α heterogeneity parameter
NIR near-infrared region D donor
SCK Smoluchowski-Collins-Kimball A acceptor
PES potential energy surface λ reorganization energy
IHB intramolecular H-bond d diameter
KIE kinetic isotopic effect M+ metal cation
IVR intramolecular vibrational-energy redistribution kb back electron transfer rate constant
VR vibrational relaxation R0 Förster distance
Me methyl kEnT energy-transfer rate constant
BET Brunauer-Emmet-Teller ksv Stern−Volmer quenching constant
CT charge transfer r distance
ET electron transfer μm micrometer
CS charge-separated P polarization
LE locally excited w/w % percent of weight per weight
PZC point of zero charge Ln lanthanides
EI electron injection
LED light-emitting diode
CB conduction band Material Structures
ICS intersystem crossing LTA Linde type A zeolites
EPR electron spin resonance LTL Linde type L zeolites
CR charge recombination H-LTL acidic LTL zeolites
HT hole transfer K-LTL potassium LTL zeolites
I⊥/I∥ depolarization ratio CD cyclodextrin
QDs quantum dots MCM41 (MCM48) Mobil Catalytic Materials of number 41
EnT energy transfer process (48)
Homo-EnT homo-energy transfer SBA15 Santa Barbara Amorphous type material
SET surface energy transfer HSA human serum albumin
p6mm 2D-hexagonal symmetry FAU Faujasite
sSMT spectroscopic single-molecule tracking MFI mordenite framework inverted
FCS fluorescence correlation spectroscopy FER ferrierite zeolites
fwhm full width at half maximum MOR mordenite zeolites
ZVIP zero valent iron particles P25 titania nanoparticles
TOC total organic carbon TiMeso crystalline mesoporous all-TiO2 structure
STEM scanning transmission electron microscopy ZSM-5 zeolite Socony Mobil-5
CIE Commission Internationale de l’Éclairage AlPO4 aluminophosphate
DSSCs dye-sensitized solar cells CTAB micelles cetyltrimethylammonium bromide mi-
PDT photodynamic therapy celles
CLSM confocal laser scanning microscopy Cal-NAM calcined-nanochannel-incorporated alu-
rGO reduced graphene oxide mina membrane
TMS-NAM trimethylsilyl-nanochannel-incorporated
Symbols and Rate Constants alumina membrane
Å Ångstrom RWY chalcogenide-based zeolite
nm nanometer PMO periodic mesoporous organosilica
ε dielectric constant Bp-PMO biphenylylene-bridged periodic mesopo-
E, E* enol, photoexcited enol rous organosilica
K, K* keto, photoexcited keto Au NPs gold nanoparticles
AN, AN* anion, photoexcited anion APO aluminophosphate
Z, Z* zwitterion, photoexcited zwitterion Lap laponite
N, N* neutral, photoexcited neutral form MCM41-SHNFs thiol-functionalized MCM41 nanofibers
T, T* tautomer, photoexcited tautomer GBZ graphene-Bi8La10O27-zeolite
S1 electronically first excited state APbX3 general perovskite formula
S0 ground-electronic singlet state MSNs mesoporous silica nanoparticles
kPT deprotonation rate constant M2PS mesoporous organosilica nanocarriers
krec recombination rate constant synthesized via the co-condensation of a
kD (k‑D) forward (reverse) diffusion-controlled rate silica precursor and a two-photon elec-
constant tron donor
ms millisecond HMSNs hollow mesoporous silica nanoparticles
ns nanosecond MOFs metal−organic frameworks
ps picosecond CMSNRs Ce6-doped mesoporous silica nanorods
fs femtosecond BSA bovine serum albumin protein
BL DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

AOT aerosol-OT [sodium bis(2-ethylhexyl) LDQ 1-[4-(4′-methyl)-2,2′-bipyridyl]-2-[4-(4′-


sulfosuccinate] N,N′-tetramethylene-2,2′-bipyridinium)]-
EMT hexagonal analog faujasite type zeolite ethene
MSUH hexagonal ordered 1D-channels meso- NB Nile blue
phase MB Methylene Blue
MCF-7 Michigan Cancer Foundation 7 Py+ pyronine
ITO indium tin oxide PyB+ pyronine B
HeLa human cervical cancer cells PyY+ pyronine Y
KB Kleinhauer-Betke cells AP aminopropyl
AO acridine orange
Molecules RhB rhodamine B
TNT trinitrotoluene Rh3B rhodamine 3B
FL fluorenone PFH+ proflavine ions
OX+ oxinine R6G rhodamine 6G
Ibu ibuprofen PBI perylene bisimide
TRITC tetramethylrhodamine isothiocyanate PTP p-terphenyl
DEAC 7-diethylaminocoumarin-3-carboxylic acid TPPy tetraphenylpyrene
butylamine ester Bp biphenylylene
CAN canthaxanthin C1 coumarin 1
MeAcr+ methylacridine PDI N,N′-bis(2,6-dimethylphenyl)-3,4,9,10-pery-
PIC cyanine (1,1′-diethyl-2-2′-cyanine iodine) lenediimide
AF acriflavine hydrochloride NI 1,8-naphthalimide
Th+ thionine DOC 3,3′-diethyloxacarbocyanine
C153 coumarin 153 SY super yellow
2-NpOH 2-naphthol IRIS3 cyanine dye
HPTS 8-hydroxypyrene-1,3,6-sulfonate (pyranine) C30 coumarin 30
6-HQ 6-hydroxyquinoline DXP N,N′-bis(2,6-dimethylphenyl)-3,4,9,10-pery-
7-HQ 7-hydroxyquinoline lenetetracarboxylic diimide
AcOH acetic acid PR149 N,N′-bis(3,5-dimethylphenyl)-3,4,9,10-pery-
HBT 2-(2′-hydroxyphenyl)benzothiazole lenetetracarboxylic diimide
HBO 2-(2′-hydroxyphenyl)benzoxazole HPDI N,N′-bis(1-hexylheptyl)-3,4,9,10-perylenete-
HBTNH2 2-[5′-N-(3-triethoxysilyl)propylurea-2′- tracarboxylic diimide
hydroxyphenyl]benzothiazole SAs salicylideneanilines
PF proflavine C-SNARF-1 (5′ and 6′)-carboxy-10-(dimethylamino)-3-
BZP benzophenone hy dr oxy s pir o[7H-b enz o[c ]xanthene-
HYZ hydrazone 7,1′(3H)-isobenzofuran]-3′-one
SAA salicylaldehyde azine FFA furfuryl alcohol
HBA-4NP (E)-2-(2′-hydroxybenzyliden)amino-4-nitro- DY-630-MI (E)-2-((E)-3-(2-(tert-butyl)-7-(diethylami-
phenol no)- 4a,8a-dihydrochromenylium-4-yl)-
3-HF 3-hydroxyflavone allylidene)-1-(6-((2-(2,5-dioxo-2,5- dihydro-
C339 coumarin 339 1H-pyrrol-1-yl)ethyl)amino)-6-oxohexyl)-
C340 coumarin 340 3,3-dimethylindoline-5-sulfonate
C343 coumarin 343 TC tetracycline
C480 coumarin 480 TXB Texbrite
PAC343 propylamide coumarin 343 PNT p-nitrotoluene
4AP 4-aminophthalimide CdS cadmium sulfide
NR Nile red DCF Diclofenac
TPC1 triphenylamine dye TSA tungstosilicic acid
YD2-o-C8 zinc-porphyrin MeO methyl orange
MPc metal phthalocyanine (Ru(bpy)3)+2 ruthenium tris-bipyridyl
CoPc cobalt phthalocyanine PMMA poly(methyl methacrylate)
PdPc palladium phthalocyanine MA methylammonium
MO-PdPc covalent bonding palladium phthalocyanine FA formamidinium
Acr+-Mes 9-mesityl-10-methylacridinium ion FITC fluorescein isothiocyanate
DPH 1,6-diphenyl-1,3,5-hexatriene RITC rhodamine B isothiocyanate
t-Ab trans-azobenzene Ru(Bphen)2PIP ruthenium(II) complex with Bphen = 4,7-
t-St trans-stilbene diphenyl-1,10-phenanthroline and PIP = 2-
An anethole phenyl-1H-imidazo[4,5-f ] [1,10]-
DCB 1,4-dicyanobenzene phenanthroline
PHNS phenosafranine BTPN naphthalene-bridged silane
(Ni(bpy)3)2+ tris(2,2′-bipyridine)nickel(II) DNT 2,4-dinitrotoluene
(Ru(bpy)32+ tris(2,2′-bipyridine)ruthenium(II) En ethylenediamine
MV methylviologen Tb(acacn) terbium-acetylacetone

BM DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Et3N triethylamine (7) Thomas, J. K. Physical Aspects of Radiation-Induced Processes


n-, t-BuNH2 n-, t-butylamine on SiO2, γ-Al2O3, Zeolites, and Clays. Chem. Rev. 2005, 105, 1683−
Tb-TTA Tb-2-thenoyltrifluoroacetone 1734.
DPA dipicolinic Acid (8) Burns, A.; Ow, H.; Wiesner, U. Fluorescent Core-Shell Silica
Yb-DBM ytterbium-1,3-diphenylpropane-1,3-dione Nanoparticles: Towards ″Lab-on-a-Particle″ Architectures for Nano-
biotechnology. Chem. Soc. Rev. 2006, 35, 1028−1042.
Eu-NTA europium-1-(2-naphthoyl)-3,3,3-trifluoroa-
(9) Schoonheydt, R. A. UV-Vis-NIR Spectroscopy and Microscopy of
cetonate Heterogeneous Catalysts. Chem. Soc. Rev. 2010, 39, 5051−5066.
TPP silane derivative porphyrin (10) Derouane, E. G.; Védrine, J. C.; Pinto, R. R.; Borges, P. M.;
PTTA-Tb3+ N,N,N1,N1-[4′-phenyl-2′2′:6,′2′-terpyridine- Costa, L.; Lemos, M. A. N. D. A.; Lemos, F.; Ribeiro, F. R. The Acidity
6,6′-diyl]bis(methylenenitrilo)tetrakis- of Zeolites: Concepts, Measurements and Relation to Catalysis: A
(acetate)-Tb3+ Review on Experimental and Theoretical Methods for the Study of
CTMR 5-carboxytetramethylrhodamine Zeolite Acidity. Catal. Rev.: Sci. Eng. 2013, 55, 454−515.
PEG−FA poly(ethylene glycol)−folic acid (11) Rimola, A.; Costa, D.; Sodupe, M.; Lambert, J.-F.; Ugliengo, P.
PDA polydopamine Silica Surface Features and Their Role in the Adsorption of
DOX doxorubicin Biomolecules: Computational Modeling and Experiments. Chem.
CTP camptothecin Rev. 2013, 113, 4216−4313.
CA enzyme carbonic anhydrase (12) Li, Y.; Yu, J. New Stories of Zeolite Structures: Their
CuAAC copper-catalyzed alkyne−azide cycloaddition Descriptions, Determinations, Predictions, and Evaluations. Chem.
Ce6 chlorine e6 Rev. 2014, 114, 7268−7316.
(13) Weckhuysen, B. M.; Yu, J. Recent Advances in Zeolite
BATA bis(alkylthio)alkene
Chemistry and Catalysis. Chem. Soc. Rev. 2015, 44, 7022−7024.
PEG polyethylene glycol (14) Mintova, S.; Jaber, M.; Valtchev, V. Nanosized Microporous
AlClPc aluminum chloride phthalocyanine Crystals: Emerging Applications. Chem. Soc. Rev. 2015, 44, 7207−
MTS 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxyme- 7233.
thoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazo- (15) Van Speybroeck, V.; Hemelsoet, K.; Joos, L.; Waroquier, M.;
lium Bell, R. G.; Catlow, C. R. A. Advances in Theory and Their
HA hydroxyapatite Application within the Field of Zeolite Chemistry. Chem. Soc. Rev.
TCPP tetrakis(4-carboxyphenyl)porphyrin 2015, 44, 7044−7111.
Z907 ruthenium dye (16) Gartzia-Rivero, L.; Banuelos, J.; Lopez-Arbeloa, I. Excitation
TGA thioglycolic acid Energy Transfer in Artificial Antennas: From Photoactive Materials to
TPDI N,N′-bis(2-(trimethylammonio)ethyl)- Molecular Assemblies. Int. Rev. Phys. Chem. 2015, 34, 515−556.
perylene-3,4,9,10-tetracarboxylic diimide (17) Chen, T.; Dong, B.; Chen, K.; Zhao, F.; Cheng, X.; Ma, C.; Lee,
SPDI N,N′-bis(3-sulfonatopropyl)perylene- S.; Zhang, P.; Kang, S.; Ha, J.; et al. Optical Super-Resolution Imaging
of Surface Reactions. Chem. Rev. 2017, 117, 7510−7537.
3,4,9,10-tetracarboxylic diimide
(18) Corma, A.; Garcia, H. Zeolite-Based Photocatalysts. Chem.
OPDI N,N′-bis(octyloxypropyl)perylene-3,4,9,10- Commun. 2004, 0, 1443−1459.
tetracarboxylic diimide (19) Yang, P.; Gai, S.; Lin, J. Functionalized Mesoporous Silica
TKMPP 1,3,6,8-tetrakis(4-methoxyphenyl)pyrene Materials for Controlled Drug Delivery. Chem. Soc. Rev. 2012, 41,
TKHPP 1,3,6,8-tetrakis(4-hydroxyphenyl)pyrene 3679−3698.
TKSPP 1,3,6,8-tetrakis[4-(3-triethoxysilylpropylami- (20) Dědeček, J.; Sobalík, Z.; Wichterlová, B. Siting and Distribution
nocarbonyloxy)-phenyl]pyrene of Framework Aluminium Atoms in Silicon-Rich Zeolites and Impact
Zn-Por-CA-PEG tetraphenylporphyrin zinc on Catalysis. Catal. Rev.: Sci. Eng. 2012, 54, 135−223.
N3 cis-bis(isothiocyanato) bis(2,2′-bipyridyl- (21) Chen, G.; Seo, J.; Yang, C.; Prasad, P. N. Nanochemistry and
4,4′-dicarboxylato ruthenium(II) Nanomaterials for Photovoltaics. Chem. Soc. Rev. 2013, 42, 8304−
N719 ditetrabutylammonium cis-bis- 8338.
(isothiocyanato)bis(2,2′-bipyridyl-4,4′- (22) Shi, J. On the Synergetic Catalytic Effect in Heterogeneous
dicarboxylato)ruthenium(II) Nanocomposite Catalysts. Chem. Rev. 2013, 113, 2139−2181.
(23) Beale, A. M.; Gao, F.; Lezcano-Gonzalez, I.; Peden, C. H. F.;
CWQ-11 carbazole-based cyanine
Szanyi, J. Recent Advances in Automotive Catalysis for NOx Emission
Control by Small-Pore Microporous Materials. Chem. Soc. Rev. 2015,
44, 7371−7405.
REFERENCES (24) Dapsens, P. Y.; Mondelli, C.; Pérez-Ramírez, J. Design of Lewis-
(1) Ramamurthy, V. Photochemistry in Organized and Constrained Acid Centres in Zeolitic Matrices for the Conversion of Renewables.
Media; VCH: New York, 1991. Chem. Soc. Rev. 2015, 44, 7025−7043.
(2) Yoon, K. B. Electron- and Charge-Transfer Reactions within (25) Mekaru, H.; Lu, J.; Tamanoi, F. Development of Mesoporous
Zeolites. Chem. Rev. 1993, 93, 321−339. Silica-Based Nanoparticles with Controlled Release Capability for
(3) Corma, A. From Microporous to Mesoporous Molecular Sieve Cancer Therapy. Adv. Drug Delivery Rev. 2015, 95, 40−49.
Materials and Their Use in Catalysis. Chem. Rev. 1997, 97, 2373− (26) Vogt, E. T. C.; Whiting, G. T.; Chowdhury, A. D.; Weckhuysen,
2420. B. M. Zeolites and Zeotypes for Oil and Gas Conversion. In Advances
(4) Hashimoto, S. Zeolite Photochemistry: Impact of Zeolites on in Catalysis, Jentoft, F. C., Ed.; Elsevier Academic Press Inc: San Diego,
Photochemistry and Feedback from Photochemistry to Zeolite 2015; Vol. 58, pp 143−314.
Science. J. Photochem. Photobiol., C 2003, 4, 19−49. (27) Martín, C.; Ziółek, M.; Douhal, A. Ultrafast and Fast Charge
(5) Calzaferri, G.; Huber, S.; Maas, H.; Minkowski, C. Host−Guest Separation Processes in Real Dye-Sensitized Solar Cells. J. Photochem.
Antenna Materials. Angew. Chem., Int. Ed. 2003, 42, 3732−3758. Photobiol., C 2016, 26, 1−30.
(6) Corma, A.; García, H. Lewis Acids: From Conventional (28) Borja, M.; Dutta, P. K. Storage of Light Energy by
Homogeneous to Green Homogeneous and Heterogeneous Catalysis. Photoelectron Transfer across a Sensitized Zeolite-Solution Interface.
Chem. Rev. 2003, 103, 4307−4366. Nature 1993, 362, 43−45.

BN DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(29) Khundkar, L. R.; Zewail, A. H. Ultrafast Molecular Reaction Reorientation Dynamics on Surface Hydrophilicity: Competing Effects
Dynamics in Real-Time: Progress over a Decade. Annu. Rev. Phys. of the Hydration Structure and Hydrogen-Bond Strength. Phys. Chem.
Chem. 1990, 41, 15−60. Chem. Phys. 2011, 13, 19911−19917.
(30) Zewail, A. H. Femtochemistry: Atomic-Scale Dynamics of the (53) Fogarty, A. C.; Duboué-Dijon, E.; Laage, D.; Thompson, W. H.
Chemical Bond. J. Phys. Chem. A 2000, 104, 5660−5694. Origins of the Non-Exponential Reorientation Dynamics of Nano-
(31) Dantus, M.; Zewail, A. H. Femtochemistry (Special Issue). confined Water. J. Chem. Phys. 2014, 141, 18C523−518C532.
Chem. Rev. 2004, 104, 1717−2124. (54) Fogarty, A. C.; Coudert, F.-X.; Boutin, A.; Laage, D.
(32) Kumpulainen, T.; Lang, B.; Rosspeintner, A.; Vauthey, E. Reorientational Dynamics of Water Confined in Zeolites. Chem-
Ultrafast Elementary Photochemical Processes of Organic Molecules PhysChem 2014, 15, 521−529.
in Liquid Solution. Chem. Rev. 2017, 117, 10826−10939. (55) Smirnov, K. S. A Molecular Dynamics Study of the Interaction
(33) Thompson, W. H. Solvation Dynamics and Proton Transfer in of Water with the External Surface of Silicalite-1. Phys. Chem. Chem.
Nanoconfined Liquids. Annu. Rev. Phys. Chem. 2011, 62, 599−619. Phys. 2017, 19, 2950−2960.
(34) Demchenko, A. P.; Tang, K.-C.; Chou, P.-T. Excited-State (56) Burris, P. C.; Laage, D.; Thompson, W. H. Simulations of the
Proton Coupled Charge Transfer Modulated by Molecular Structure Infrared, Raman, and 2D-IR Photon Echo Spectra of Water in
and Media Polarization. Chem. Soc. Rev. 2013, 42, 1379−1408. Nanoscale Silica Pores. J. Chem. Phys. 2016, 144, 194709−194720.
(35) Bagchi, B. Dynamics of Solvation and Charge Transfer (57) Cimas, À .; Tielens, F.; Sulpizi, M.; Gaigeot, M.-P.; Costa, D.
Reactions in Dipolar Liquids. Annu. Rev. Phys. Chem. 1989, 40, The Amorphous Silica−Liquid Water Interface Studied by Ab Initio
115−141. Molecular Dynamics (AIMD): Local Organization in Global Disorder.
(36) Glasbeek, M.; Zhang, H. Femtosecond Studies of Solvation and J. Phys.: Condens. Matter 2014, 26, 244106−244115.
Intramolecular Configurational Dynamics of Fluorophores in Liquid (58) Gierada, M.; Petit, I.; Handzlik, J.; Tielens, F. Hydration in Silica
Solution. Chem. Rev. 2004, 104, 1929−1954. Based Mesoporous Materials: A DFT Model. Phys. Chem. Chem. Phys.
(37) Kosower, E. M.; Huppert, D. Excited State Electron and Proton 2016, 18, 32962−32972.
Transfers. Annu. Rev. Phys. Chem. 1986, 37, 127−156. (59) Ding, F.; Hu, Z.; Zhong, Q.; Manfred, K.; Gattass, R. R.;
(38) Tomin, V. I.; Demchenko, A. P.; Chou, P.-T. Thermodynamic Brindza, M. R.; Fourkas, J. T.; Walker, R. A.; Weeks, J. D. Interfacial
Vs. Kinetic Control of Excited-State Proton Transfer Reactions. J. Organization of Acetonitrile: Simulation and Experiment. J. Phys.
Photochem. Photobiol., C 2015, 22, 1−18. Chem. C 2010, 114, 17651−17659.
(39) Mohammed, O. F.; Pines, D.; Dreyer, J.; Pines, E.; Nibbering, E. (60) Liu, S.; Hu, Z.; Weeks, J. D.; Fourkas, J. T. Structure of Liquid
T. J. Sequential Proton Transfer through Water Bridges in Acid-Base Propionitrile at Interfaces. 1. Molecular Dynamics Simulations. J. Phys.
Reactions. Science 2005, 310, 83−86. Chem. C 2012, 116, 4012−4018.
(40) Rini, M.; Magnes, B.-Z.; Pines, E.; Nibbering, E. T. Real-Time (61) Roy, D.; Liu, S.; Woods, B. L.; Siler, A. R.; Fourkas, J. T.; Weeks,
Observation of Bimodal Proton Transfer in Acid-Base Pairs in Water. J. D.; Walker, R. A. Nonpolar Adsorption at the Silica/Methanol
Science 2003, 301, 349−352. Interface: Surface Mediated Polarity and Solvent Density across a
(41) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Structural Changes Strongly Associating Solid/Liquid Boundary. J. Phys. Chem. C 2013,
Accompanying Intramolecular Electron Transfer: Focus on Twisted 117, 27052−27061.
Intramolecular Charge-Transfer States and Structures. Chem. Rev. (62) Coasne, B.; Fourkas, J. T. Structure and Dynamics of Benzene
2003, 103, 3899−4032. Confined in Silica Nanopores. J. Phys. Chem. C 2011, 115, 15471−
(42) Maroncelli, M.; MacInnis, J.; Fleming, G. R. Polar Solvent 15479.
Dynamics and Electron-Transfer Reactions. Science 1989, 243, 1674− (63) Wells, R. H.; Thompson, W. H. What Determines the Location
1681. of a Small Solute in a Nanoconfined Liquid? J. Phys. Chem. B 2015,
(43) Tamai, N.; Miyasaka, H. Ultrafast Dynamics of Photochromic 119, 12446−12454.
Systems. Chem. Rev. 2000, 100, 1875−1890. (64) Vartia, A. A.; Thompson, W. H. Solvation and Spectra of a
(44) Douhal, A. Ultrafast Guest Dynamics in Cyclodextrin Charge Transfer Solute in Ethanol Confined within Nanoscale Silica
Nanocavities. Chem. Rev. 2004, 104, 1955−1976. Pores. J. Phys. Chem. B 2012, 116, 5414−5424.
(45) Douhal, A. Fast and Ultrafast Dynamics in Cyclodextrin (65) Das, A.; Chakrabarti, J. Microscopic Mechanisms of Confine-
Nanostructures. In Cyclodextrin Materials Photochemistry, Photophysics ment-Induced Slow Solvation. J. Phys. Chem. A 2013, 117, 10571−
and Photobiology, Douhal, A., Ed.; Elsevier: Amsterdam, 2006; Vol. 1, 10575.
Chapter 8, pp 181−201. (66) Harvey, J. A.; Thompson, W. H. Thermodynamic Driving
(46) Nandi, N.; Bhattacharyya, K.; Bagchi, B. Dielectric Relaxation Forces for Dye Molecule Position and Orientation in Nanoconfined
and Solvation Dynamics of Water in Complex Chemical and Biological Solvents. J. Phys. Chem. B 2015, 119, 9150−9159.
Systems. Chem. Rev. 2000, 100, 2013−2046. (67) Elola, M. D.; Rodriguez, J.; Laria, D. Liquid Methanol Confined
(47) Bagchi, B. Water Dynamics in the Hydration Layer around within Functionalized Silica Nanopores. 2. Solvation Dynamics of
Proteins and Micelles. Chem. Rev. 2005, 105, 3197−3219. Coumarin 153. J. Phys. Chem. B 2011, 115, 12859−12867.
(48) Janssen, K. P. F.; De Cremer, G.; Neely, R. K.; Kubarev, A. V.; (68) Cazade, P. A.; Bordat, P.; Baraille, I.; Brown, R.; Smith, W.;
Van Loon, J.; Martens, J. A.; De Vos, D. E.; Roeffaers, M. B. J.; Todorov, I. T. DL_Poly_2 Adaptations for Solvation Studies. Mol.
Hofkens, J. Single Molecule Methods for the Study of Catalysis: From Simul. 2011, 37, 43−52.
Enzymes to Heterogeneous Catalysts. Chem. Soc. Rev. 2014, 43, 990− (69) Bordat, P.; Cazade, P.-A.; Baraille, I.; Brown, R. Host and
1006. Adsorbate Dynamics in Silicates with Flexible Frameworks: Empirical
(49) Martín, C.; Gil, M.; Cohen, B.; Douhal, A. Ultrafast Force Field Simulation of Water in Silicalite. J. Chem. Phys. 2010, 132,
Photodynamics of Drugs in Nanocavities: Cyclodextrins and Human 094501−094508.
Serum Albumin Protein. Langmuir 2012, 28, 6746−6759. (70) Rodriguez, J.; Elola, M. D. Molecular Dynamics Simulations of
(50) Zhu, W.; Wang, C.; Wang, H.; Li, G. Theory and Simulation of Ibuprofen Release from pH-Gated Silica Nanochannels. J. Phys. Chem.
Diffusion-Adsorption into a Molecularly Imprinted Mesoporous Film B 2015, 119, 8868−8878.
and Its Nanostructured Counterparts. Experimental Application for (71) Fois, E.; Tabacchi, G.; Calzaferri, G. Interactions, Behavior, and
Trace Explosive Detection. RSC Adv. 2014, 4, 40676−40685. Stability of Fluorenone inside Zeolite Nanochannels. J. Phys. Chem. C
(51) Laage, D.; Thompson, W. H. Reorientation Dynamics of 2010, 114, 10572−10579.
Nanoconfined Water: Power-Law Decay, Hydrogen-Bond Jumps, and (72) Fois, E.; Tabacchi, G.; Calzaferri, G. Orientation and Order of
Test of a Two-State Model. J. Chem. Phys. 2012, 136, 044513−004519. Xanthene Dyes in the One-Dimensional Channels of Zeolite L:
(52) Stirnemann, G.; Castrillon, S. R.-V.; Hynes, J. T.; Rossky, P. J.; Bridging the Gap between Experimental Data and Molecular Behavior.
Debenedetti, P. G.; Laage, D. Non-Monotonic Dependence of Water J. Phys. Chem. C 2012, 116, 16784−16799.

BO DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(73) Zhou, X.; Wesolowski, T. A.; Tabacchi, G.; Fois, E.; Calzaferri, (88) Lutkouskaya, K.; Calzaferri, G. Transfer of Electronic Excitation
G.; Devaux, A. First-Principles Simulation of the Absorption Bands of Energy between Randomly Mixed Dye Molecules in the Channels of
Fluorenone in Zeolite L. Phys. Chem. Chem. Phys. 2013, 15, 159−167. Zeolite L. J. Phys. Chem. B 2006, 110, 5633−5638.
(74) Delle Piane, M.; Vaccari, S.; Corno, M.; Ugliengo, P. Silica- (89) Berberan-Santos, M. N.; Bodunov, E. N.; Valeur, B.
Based Materials as Drug Adsorbents: First Principle Investigation on Mathematical Functions for the Analysis of Luminescence Decays
the Role of Water Microsolvation on Ibuprofen Adsorption. J. Phys. with Underlying Distributions 1. Kohlrausch Decay Function
Chem. A 2014, 118, 5801−5807. (Stretched Exponential). Chem. Phys. 2005, 315, 171−182.
(75) de Queiroz, T. B.; Botelho, M. B. S.; Fernández-Hernández, J. (90) Lee, K. C. B.; Siegel, J.; Webb, S. E. D.; Lévêque-Fort, S.; Cole,
M.; Eckert, H.; Albuquerque, R. Q.; de Camargo, A. S. S. New M. J.; Jones, R.; Dowling, K.; Lever, M. J.; French, P. M. W.
Luminescent Host−Guest System Based on an Iridium(III) Complex: Application of the Stretched Exponential Function to Fluorescence
Design, Synthesis, and Theoretical−Experimental Spectroscopic Lifetime Imaging. Biophys. J. 2001, 81, 1265−1274.
Characterization. J. Phys. Chem. C 2013, 117, 2966−2975. (91) Marquis, S.; Moissette, A.; Brémard, C. Incorporation of
(76) Chandra Mohan, S.; Bhattacharjee, D.; Chandra Deka, R.; Anthracene into Zeolites: Confinement Effect on the Recombination
Jothivenkatachalam, K. Combined Experimental and Theoretical Rate of Photoinduced Radical Cation-Electron Pair. ChemPhysChem
Investigations on the Encapsulation of Nickel(II)tet-a Complex in 2006, 7, 1525−1534.
Zeolite Y and Its Photocatalytic Activity. RSC Adv. 2016, 6, 71214− (92) Clifford, J. N.; Palomares, E.; Nazeeruddin, M. K.; Grätzel, M.;
71222. Nelson, J.; Li, X.; Long, N. J.; Durrant, J. R. Molecular Control of
(77) Van Speybroeck, V.; Hemelsoet, K.; De Wispelaere, K.; Qian, Recombination Dynamics in Dye-Sensitized Nanocrystalline TiO2
Q.; Van der Mynsbrugge, J.; De Sterck, B.; Weckhuysen, B. M.; Films: Free Energy vs Distance Dependence. J. Am. Chem. Soc.
Waroquier, M. Mechanistic Studies on Chabazite-Type Methanol-to- 2004, 126, 5225−5233.
Olefin Catalysts: Insights from Time-Resolved UV/Vis Microspectro- (93) Viani, L.; Minoia, A.; Cornil, J.; Beljonne, D.; Egelhaaf, H.-J.;
scopy Combined with Theoretical Simulations. ChemCatChem 2013, Gierschner, J. Resonant Energy Transport in Dye-Filled Monolithic
5, 173−184. Crystals of Zeolite L: Modeling of Inhomogeneity. J. Phys. Chem. C
(78) Hemelsoet, K.; Qian, Q.; De Meyer, T.; De Wispelaere, K.; De 2016, 120, 27192−27199.
Sterck, B.; Weckhuysen, B. M.; Waroquier, M.; Van Speybroeck, V. (94) Fois, E.; Tabacchi, G.; Devaux, A.; Belser, P.; Brühwiler, D.;
Identification of Intermediates in Zeolite-Catalyzed Reactions by in Calzaferri, G. Host−Guest Interactions and Orientation of Dyes in the
Situ UV/Vis Microspectroscopy and a Complementary Set of
One-Dimensional Channels of Zeolite L. Langmuir 2013, 29, 9188−
Molecular Simulations. Chem. - Eur. J. 2013, 19, 16595−16606.
9198.
(79) Insuwan, W.; Rangsriwatananon, K.; Meeprasert, J.;
(95) Insuwan, W.; Rangsriwatananon, K.; Meeprasert, J.;
Namuangruk, S.; Surakhot, Y.; Kungwan, N.; Jungsuttiwong, S.
Namuangruk, S.; Surakhot, Y.; Kungwan, N.; Jungsuttiwong, S.
Combined Experimental and Theoretical Investigation on Photo-
Combined Experimental and Theoretical Investigation on Fluores-
physical Properties of Trans-Azobenzene Confined in LTL Zeolite:
cence Resonance Energy Transfer of Dye Loaded on LTL Zeolite.
Effect of Cis-Isomer Forming. Microporous Mesoporous Mater. 2014,
Microporous Mesoporous Mater. 2017, 241, 372−382.
197, 348−357.
(96) Eigen, M. Proton Transfer, Acid-Base Catalysis, and Enzymatic
(80) Dong, C.; Li, X.; Qi, J. Probing the Electronic and Optical
Properties of Silica-Coated Quantum Dots with First-Principles Hydrolysis. Part I: Elementary Processes. Angew. Chem., Int. Ed. Engl.
Calculations. Phys. Chem. Chem. Phys. 2011, 13, 14476−14480. 1964, 3, 1−19.
(81) Pedone, A.; Prampolini, G.; Monti, S.; Barone, V. Absorption (97) Kasha, M. Proton-Transfer Spectroscopy. Perturbation of the
and Emission Spectra of Fluorescent Silica Nanoparticles from TD- Tautomerization Potential. J. Chem. Soc., Faraday Trans. 2 1986, 82,
DFT/MM/PCM Calculations. Phys. Chem. Chem. Phys. 2011, 13, 2379−2392.
16689−16697. (98) Arnaut, L. G.; Formosinho, S. J. Excited-State Proton Transfer
(82) Pedone, A.; Prampolini, G.; Monti, S.; Barone, V. Realistic Reactions I. Fundamentals and Intermolecular Reactions. J. Photochem.
Modeling of Fluorescent Dye-Doped Silica Nanoparticles: A Step Photobiol., A 1993, 75, 1−20.
toward the Understanding of Their Enhanced Photophysical Proper- (99) Formosinho, S. J.; Arnaut, L. G. Excited-State Proton Transfer
ties. Chem. Mater. 2011, 23, 5016−5023. Reactions II. Intramolecular Reactions. J. Photochem. Photobiol., A
(83) Pedone, A.; Bloino, J.; Barone, V. Role of Host−Guest 1993, 75, 21−48.
Interactions in Tuning the Optical Properties of Coumarin Derivatives (100) Douhal, A.; Kim, S. K.; Zewail, A. H. Femtosecond Molecular
Incorporated in MCM-41: A TD-DFT Investigation. J. Phys. Chem. C Dynamics of Tautomerization in Model Base Pairs. Nature 1995, 378,
2012, 116, 17807−17818. 260−263.
(84) Pedone, A.; Gambuzzi, E.; Barone, V.; Bonacchi, S.; Genovese, (101) Douhal, A.; Lahmani, F.; Zewail, A. H. Proton-Transfer
D.; Rampazzo, E.; Prodi, L.; Montalti, M. Understanding the Reaction Dynamics. Chem. Phys. 1996, 207, 477−498.
Photophysical Properties of Coumarin-Based Pluronic-Silica (PluS) (102) Mohammed, O. F.; Pines, D.; Nibbering, E. T. J.; Pines, E.
Nanoparticles by Means of Time-Resolved Emission Spectroscopy and Base-Induced Solvent Switches in Acid−Base Reactions. Angew. Chem.,
Accurate TD-DFT/Stochastic Calculations. Phys. Chem. Chem. Phys. Int. Ed. 2007, 46, 1458−1461.
2013, 15, 12360−12372. (103) Pérez-Lustres, J. L.; Rodriguez-Prieto, F.; Mosquera, M.;
(85) Barone, V.; Biczysko, M.; Bloino, J.; Carta, L.; Pedone, A. Senyushkina, T. A.; Ernsting, N. P.; Kovalenko, S. A. Ultrafast Proton
Environmental and Dynamical Effects on the Optical Properties of Transfer to Solvent: Molecularity and Intermediates from Solvation-
Molecular Systems by Time-Independent and Time-Dependent and Diffusion-Controlled Regimes. J. Am. Chem. Soc. 2007, 129,
Approaches: Coumarin Derivatives as Test Cases. Comput. Theor. 5408−5418.
Chem. 2014, 1037, 35−48. (104) Rini, M.; Magnes, B.-Z.; Pines, E.; Nibbering, E. T. J. Real-
(86) Gao, Y.; Xu, D.; Kispert, L. D. Hydrogen Bond Formation Time Observation of Bimodal Proton Transfer in Acid-Base Pairs in
between the Carotenoid Canthaxanthin and the Silanol Group on Water. Science 2003, 301, 349−352.
MCM-41 Surface. J. Phys. Chem. B 2015, 119, 10488−10495. (105) Tomin, V. I. Proton Transfer Reactions in the Excited
(87) Gao, Y.; Chen, H.; Tay-Agbozo, S.; Kispert, L. D. Photo- Electronic State. In Hydrogen Bonding and Transfer in the Excited State;
Induced Electron Transfer of Carotenoids in Mesoporous Sieves John Wiley & Sons Ltd: Hoboken, 2010; pp 463−523.
(MCM-41) and Surface Modified MCM-41: The Role of Hydrogen (106) Elsässer, T.; Van den Akker, H. Ultrafast Hydrogen Bonding
Bonds on the Electron Transfer. J. Photochem. Photobiol., A 2017, 341, Dynamics and Proton Transfer Processes in the Condensed Phase;
1−11. Springer Science & Business Media: New York, 2003; Vol. 23.

BP DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(107) Hynes, J. T.; Klinman, J. P.; Limbach, H.-H.; Schowen, R. L. (125) Ireland, J. F.; Wyatt, P. A. H. Acid-Base Properties of
Hydrogen-Transfer Reactions; Wiley-VCH: Weinheim, Germany, 2007; Electronically Excited States of Organic Molecules. In Advances in
Vol. 1−4. Physical Organic Chemistry, Gold, V., Ed.; Academic Press: London,
(108) Wu, J.; Liu, W.; Ge, J.; Zhang, H.; Wang, P. New Sensing 1976; Vol. 12, pp 131−221.
Mechanisms for Design of Fluorescent Chemosensors Emerging in (126) Spry, D. B.; Goun, A.; Fayer, M. D. Deprotonation Dynamics
Recent Years. Chem. Soc. Rev. 2011, 40, 3483−3495. and Stokes Shift of Pyranine (HPTS). J. Phys. Chem. A 2007, 111,
(109) Zhao, J.; Ji, S.; Chen, Y.; Guo, H.; Yang, P. Excited State 230−237.
Intramolecular Proton Transfer (ESIPT): From Principal Photo- (127) Cohen, B.; Huppert, D.; Agmon, N. Diffusion-Limited Acid−
physics to the Development of New Chromophores and Applications Base Nonexponential Dynamics. J. Phys. Chem. A 2001, 105, 7165−
in Fluorescent Molecular Probes and Luminescent Materials. Phys. 7173.
Chem. Chem. Phys. 2012, 14, 8803−8817. (128) Weller, A. Protolytic Reactions of Excited Hydroxy
(110) Costela, A.; Muñoz, J. M.; Douhal, A.; Figuera, J. M.; Acuña, A. Compounds. Z. Phys. Chem. 1958, 17, 224−225.
U. Experimental Test of a Four-Level Kinetic Model for Excited-State (129) Weller, A. Fast Reactions of Excited Molecules. In Progress in
Intramolecular Proton Transfer Dye Lasers. Appl. Phys. B: Photophys. Reaction Kinetics, Porter, G., Ed. Pergamon: New York, 1961; Vol. 1,
Laser Chem. 1989, 49, 545−552. pp 197−214.
(111) Sastre, R.; Costela, A. Polymeric Solid-State Dye Lasers. Adv. (130) Eigen, M.; Kruse, W.; Maass, G.; De Maeyer, L. Rate
Mater. 1995, 7, 198−202. Constants of Protolytic Reactions in Aqueous Solution. In Progress in
(112) Costela, A.; Garcia-Moreno, I.; Figuera, J. M.; Amat-Guerri, F.; Reaction Kinetics, Porter, G., Ed. Pergamon: New York, 1964; Vol. 2,
Sastre, R. Solid-State Dye Lasers Based on Polymers Incorporating pp 286−318
Covalently Bonded Modified Rhodamine 6G. Appl. Phys. Lett. 1996, (131) Cohen, B.; Huppert, D.; Agmon, N. Non-Exponential
68, 593−595. Smoluchowski Dynamics in Fast Acid−Base Reaction. J. Am. Chem.
(113) Martin, E.; Weigand, R.; Pardo, A. Solvent Dependence of the Soc. 2000, 122, 9838−9839.
Inhibition of Intramolecular Charge-Transfer in N-Substituted 1,8- (132) Pines, E.; Huppert, D.; Agmon, N. Geminate Recombination
Naphthalimide Derivatives as Dye Lasers. J. Lumin. 1996, 68, 157− in Excited-State Proton-Transfer Reactions: Numerical Solution of the
164. Debye−Smoluchowski Equation with Backreaction and Comparison
(114) Jones, G.; Jackson, W. R.; Choi, C. Y.; Bergmark, W. R. Solvent
with Experimental Results. J. Chem. Phys. 1988, 88, 5620−5630.
Effects on Emission Yield and Lifetime for Coumarin Laser Dyes. (133) Borgis, D.; Hynes, J. T. Curve Crossing Formulation for
Requirements for a Rotatory Decay Mechanism. J. Phys. Chem. 1985,
Proton Transfer Reactions in Solution. J. Phys. Chem. 1996, 100,
89, 294−300.
1118−1128.
(115) Fluegge, A. P.; Waiblinger, F.; Stein, M.; Keck, J.; Kramer, H.;
(134) Smoluchowski, M. Versuch Einer Mathematischen Theorie
Fischer, P.; Wood, M. G.; DeBellis, A. D.; Ravichandran, R.; Leppard,
Der Koagulationskinetik Kolloider Lösungen. Z. Phys. Chem. 1918, 92,
D. Probing the Intramolecular Hydrogen Bond of 2-(2-
129−168.
Hydroxyphenyl)Benzotriazoles in Polar Environment: A Photo-
(135) Krissinel, E. B.; Agmon, N. Spherical Symmetric Diffusion
physical Study of UV Absorber Efficiency. J. Phys. Chem. A 2007,
Problem. J. Comput. Chem. 1996, 17, 1085−1098.
111, 9733−9744.
(136) Szabo, A. Theory of Diffusion-Influenced Fluorescence
(116) Lapinski, L.; Nowak, M. J.; Nowacki, J.; Rode, M. F.;
Quenching. J. Phys. Chem. 1989, 93, 6929−6939.
Sobolewski, A. L. A Bistable Molecular Switch Driven by Photo-
(137) Rice, S. A. Diffusion-Limited Reactions; Elsevier, 1985; Vol. 25.
induced Hydrogen-Atom Transfer. ChemPhysChem 2009, 10, 2290−
(138) Solntsev, K. M.; Clower, C. E.; Tolbert, L. M.; Huppert, D. 6-
2295.
Hydroxyquinoline-N-Oxides: A New Class of “Super” Photoacids1. J.
(117) Lim, S.-J.; Seo, J.; Park, S. Y. Photochromic Switching of
Am. Chem. Soc. 2005, 127, 8534−8544.
Excited-State Intramolecular Proton-Transfer (ESIPT) Fluorescence:
(139) Kwon, O.-H.; Lee, Y.-S.; Yoo, B.; Jang, D.-J. Excited-State
A Unique Route to High-Contrast Memory Switching and Non-
Triple Proton Transfer of 7-Hydroxyquinoline Along a Hydrogen-
destructive Readout. J. Am. Chem. Soc. 2006, 128, 14542−14547.
(118) Mutai, T.; Tomoda, H.; Ohkawa, T.; Yabe, Y.; Araki, K. Bonded Alcohol Chain: Vibrationally Assisted Proton Tunneling.
Switching of Polymorph-Dependent Esipt Luminescence of an Angew. Chem., Int. Ed. 2006, 45, 415−419.
Imidazo[1,2-a]Pyridine Derivative. Angew. Chem. 2008, 120, 9664− (140) Park, S.-Y.; Kwon, O.-H.; Lee, Y.-S.; Jang, D.-J. Triplet-State
9666. Acid−Base Reactions of 1-Methyl-7-Oxyquinolinium in Water. J. Phys.
(119) Sliwa, M.; Létard, S.; Malfant, I.; Nierlich, M.; Lacroix, P. G.; Chem. A 2009, 113, 10589−10592.
Asahi, T.; Masuhara, H.; Yu, P.; Nakatani, K. Design, Synthesis, (141) Agmon, N.; Huppert, D.; Masad, A.; Pines, E. Excited-State
Structural and Nonlinear Optical Properties of Photochromic Crystals: Proton Transfer to Methanol-Water Mixtures. J. Phys. Chem. 1991, 95,
Toward Reversible Molecular Switches. Chem. Mater. 2005, 17, 4727− 10407−10413.
4735. (142) Klöpffer, W. Intramolecular Proton Transfer in Electronically
(120) Kwon, J. E.; Park, S. Y. Advanced Organic Optoelectronic Excited Molecules. In Advances in Photochemistry; John Wiley & Sons,
Materials: Harnessing Excited-State Intramolecular Proton Transfer Inc.: Hoboken, NJ, 2007; Vol. 10, pp 311−358.
(ESIPT) Process. Adv. Mater. 2011, 23, 3615−3642. (143) Gil, M.; Martin, C.; Douhal, A. Femtosecond Dynamics and
(121) Park, S.; Seo, J.; Kim, S. H.; Park, S. Y. Tetraphenylimidazole- Photoconversion of a H-Bonded Dye within Mesoporous Silicate
Based Excited-State Intramolecular Proton-Transfer Molecules for Materials. J. Phys. Chem. C 2011, 115, 14687−14697.
Highly Efficient Blue Electroluminescence. Adv. Funct. Mater. 2008, (144) Gil, M.; Martin, C.; Organero, J. A.; Navarro, M. T.; Corma,
18, 726−731. A.; Douhal, A. Interrogating Confined Proton-Transfer Reaction
(122) Park, S.; Kwon, J.; Kim, S.; Seo, J.; Chung, K.; Park, S.; Jang, Dynamics within Mesoporous Nanotubes. J. Phys. Chem. C 2010, 114,
D.; Medina, B.; Gierschner, J.; Park, S. Y. A White-Light-Emitting 6311−6317.
Molecule: Frustrated Energy Transfer between Constituent Emitting (145) Abou-Zied, O. K.; Jimenez, R.; Thompson, E. H. Z.; Millar, D.
Centers. J. Am. Chem. Soc. 2009, 131, 14043−14049. P.; Romesberg, F. E. Solvent-Dependent Photoinduced Tautomeriza-
(123) Tolbert, L. M.; Solntsev, K. M. Excited-State Proton Transfer: tion of 2-(2′-Hydroxyphenyl)Benzoxazole. J. Phys. Chem. A 2002, 106,
From Constrained Systems to “Super” Photoacids to Superfast Proton 3665−3672.
Transfer†. Acc. Chem. Res. 2002, 35, 19−27. (146) Woolfe, G. J.; Melzig, M.; Schneider, S.; Dörr, F. The Role of
(124) Agmon, N. Elementary Steps in Excited-State Proton Transfer. Tautomeric and Rotameric Species in the Photophysics of 2-(2′-
J. Phys. Chem. A 2005, 109, 13−35. Hydroxyphenyl)Benzoxazole. Chem. Phys. 1983, 77, 213−221.

BQ DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(147) Al-Lawatia, N.; Husband, J.; Steinbrecher, T.; Abou-Zied, O. K. (167) Lochbrunner, S.; Wurzer, A. J.; Riedle, E. Microscopic
Tautomerism in 7-Hydroxyquinoline: A Combined Experimental and Mechanism of Ultrafast Excited-State Intramolecular Proton Transfer:
Theoretical Study in Water. J. Phys. Chem. A 2011, 115, 4195−4201. A 30-fs Study of 2-(2′-Hydroxyphenyl)Benzothiazole. J. Phys. Chem. A
(148) Abou-Zied, O. K.; Husband, J.; Al-Lawatia, N.; Steinbrecher, T. 2003, 107, 10580−10590.
B. Ground State Spectroscopy of Hydroxyquinolines: Evidence for the (168) Kim, C. H.; Joo, T. Coherent Excited State Intramolecular
Formation of Protonated Species in Water-Rich Dioxane-Water Proton Transfer Probed by Time-Resolved Fluorescence. Phys. Chem.
Mixtures. Phys. Chem. Chem. Phys. 2014, 16, 61−70. Chem. Phys. 2009, 11, 10266−10269.
(149) Barbara, P. F.; Walsh, P. K.; Brus, L. E. Picosecond Kinetic and (169) Rodríguez-Córdoba, W.; Zugazagoitia, J. S.; Collado-Fregoso,
Vibrationally Resolved Spectroscopic Studies of Intramolecular E.; Peon, J. Excited State Intramolecular Proton Transfer in Schiff
Excited-State Hydrogen Atom Transfer. J. Phys. Chem. 1989, 93, Bases. Decay of the Locally Excited Enol State Observed by
29−34. Femtosecond Resolved Fluorescence. J. Phys. Chem. A 2007, 111,
(150) Naundorf, H.; Organero, J. A.; Douhal, A.; Kühn, O. Potential 6241−6247.
Energy Surface for the Proton Transfer in 8-Hydroxyimidazo[1,2- (170) Kühn, O.; González, L. Laser-Driven Ultrafast Hydrogen
a]Pyridine. J. Chem. Phys. 1999, 110, 11286−11293. Transfer Dynamics. Hydrogen-Transfer Reactions; Wiley-VCH: Wein-
(151) Batista, V. S.; Brumer, P. Coherent Control in the Presence of heim, Germany, 2006; Vol. 1, pp 79−10310.1002/
Intrinsic Decoherence: Proton Transfer in Large Molecular Systems. 9783527611546.ch4.
Phys. Rev. Lett. 2002, 89, 143201−143204. (171) Hsieh, C.-C.; Jiang, C.-M.; Chou, P.-T. Recent Experimental
(152) Douhal, A. A Quick Look at Hydrogen Bonds. Science 1997, Advances on Excited-State Intramolecular Proton Coupled Electron
276, 221−222. Transfer Reaction. Acc. Chem. Res. 2010, 43, 1364−1374.
(153) Matanović, I.; Došlić, N.; Mihalić, Z. Exploring the Potential (172) Hammes-Schiffer, S.; Stuchebrukhov, A. A. Theory of Coupled
Energy Surface for Proton Transfer in Acetylacetone. Chem. Phys. Electron and Proton Transfer Reactions. Chem. Rev. 2010, 110, 6939−
2004, 306, 201−207. 6960.
(154) Dolati, F.; Tayyari, S. F.; Vakili, M.; Wang, Y. A. Proton (173) Huynh, M. H. V.; Meyer, T. J. Proton-Coupled Electron
Transfer in Acetylacetone and Its α-Halo Derivatives. Phys. Chem. Transfer. Chem. Rev. 2007, 107, 5004−5064.
Chem. Phys. 2016, 18, 344−350. (174) Genosar, L.; Cohen, B.; Huppert, D. Ultrafast Direct
(155) Gertitschke, P. L.; Kiprof, P.; Manz, J. A Dynamical ‘‘White Photoacid−Base Reaction. J. Phys. Chem. A 2000, 104, 6689−6698.
Spot’’ on the Potential Energy Surface: The Close Interaction Region (175) Hynes, J. T.; Tran-Thi, T.-H.; Granucci, G. Intermolecular
of the Collinear Hydrogen Transfer Reaction F+DBr→FD+Br. J. Photochemical Proton Transfer in Solution: New Insights and
Chem. Phys. 1987, 87, 941−952. Perspectives. J. Photochem. Photobiol., A 2002, 154, 3−11.
(156) Herek, J. L.; Pedersen, S.; Bañares, L.; Zewail, A. H. (176) Leiderman, P.; Genosar, L.; Huppert, D. Excited-State Proton
Femtosecond Real-Time Probing of Reactions. IX. Hydrogen-Atom Transfer: Indication of Three Steps in the Dissociation and
Transfer. J. Chem. Phys. 1992, 97, 9046−9061. Recombination Process. J. Phys. Chem. A 2005, 109, 5965−5977.
(157) Zhong, D. P.; Douhal, A.; Zewail, A. H. Femtosecond Studies (177) Liu, W.; Han, F.; Smith, C.; Fang, C. Ultrafast Conformational
of Protein-Ligand Hydrophobic Binding and Dynamics: Human Dynamics of Pyranine During Excited State Proton Transfer in
Serum Albumin. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 14056−14061. Aqueous Solution Revealed by Femtosecond Stimulated Raman
(158) Lu, C.; Hsieh, R. R.; Lee, I. R.; Cheng, P.-Y. Ultrafast Spectroscopy. J. Phys. Chem. B 2012, 116, 10535−10550.
Dynamics of Gas-Phase Excited-State Intramolecular Proton Transfer (178) Siwick, B. J.; Bakker, H. J. On the Role of Water in
in 1-Hydroxy-2-Acetonaphthone. Chem. Phys. Lett. 1999, 310, 103− Intermolecular Proton-Transfer Reactions. J. Am. Chem. Soc. 2007,
110. 129, 13412−13420.
(159) Organero, J. A.; Tormo, L.; Douhal, A. Caging Ultrafast Proton (179) Lee, Y.-S.; Yu, H.; Kwon, O.-H.; Jang, D.-J. Photo-Induced
Transfer and Twisting Motion of 1-Hydroxy-2-Acetonaphthone. Proton-Transfer Cycle of 2-Naphthol in Faujasite Zeolitic Nano-
Chem. Phys. Lett. 2002, 363, 409−414. cavities. Phys. Chem. Chem. Phys. 2008, 10, 153−158.
(160) Douhal, A.; Lahmani, F.; Zehnacker-Rentien, A. Excited-State (180) Hineman, M. F.; Brucker, G. A.; Kelley, D. F.; Bernstein, E. R.
Intramolecular Proton Transfer in Jet-Cooled 1-Hydroxy-2-Aceto- Excited-State Proton Transfer in 1-Naphthol/Ammonia Clusters. J.
naphthone. Chem. Phys. 1993, 178, 493−504. Chem. Phys. 1992, 97, 3341−3347.
(161) Alarcos, N.; Gutierrez, M.; Liras, M.; Sanchez, F.; Douhal, A. (181) Mandal, D.; Pal, S. K.; Bhattacharyya, K. Excited-State Proton
An Abnormally Slow Proton Transfer Reaction in a Simple HBO Transfer of 1-Naphthol in Micelles. J. Phys. Chem. A 1998, 102, 9710−
Derivative Due to Ultrafast Intramolecular-Charge Transfer Events. 9714.
Phys. Chem. Chem. Phys. 2015, 17, 16257−16269. (182) Park, H.-R.; Mayer, B.; Wolschann, P.; Koehler, G. Excited-
(162) Alarcos, N.; Gutierrez, M.; Liras, M.; Sanchez, F.; Moreno, M.; State Proton Transfer of 2-Naphthol Inclusion Complexes with
Douhal, A. Direct Observation of Breaking of the Intramolecular H- Cyclodextrins. J. Phys. Chem. 1994, 98, 6158−6166.
Bond, and Slowing Down of the Proton Motion and Tuning Its (183) Laws, W. R.; Brand, L. Analysis of Two-State Excited-State
Mechanism in an HBO Derivative. Phys. Chem. Chem. Phys. 2015, 17, Reactions. The Fluorescence Decay of 2-Naphthol. J. Phys. Chem.
14569−14581. 1979, 83, 795−802.
(163) Gutierrez, M.; Alarcos, N.; Liras, M.; Sánchez, F.; Douhal, A. (184) Cohen, B.; Leiderman, P.; Huppert, D. Unusual Temperature
Switching to a Reversible Proton Motion in a Charge-Transferred Dye. Dependence of Proton Transfer. 2. Excited-State Proton Transfer from
J. Phys. Chem. B 2015, 119, 552−562. Photoacids to Water. J. Phys. Chem. A 2002, 106, 11115−11122.
(164) Kim, C. H.; Park, J.; Seo, J.; Park, S. Y.; Joo, T. Excited State (185) García-Ochoa, I.; Díez López, M. A.; Viñas, M. H.; Santos, L.;
Intramolecular Proton Transfer and Charge Transfer Dynamics of a 2- Martínez Ataz, E.; Sánchez, F.; Douhal, A. Hydrogen-Bonding
(2′-Hydroxyphenyl)Benzoxazole Derivative in Solution. J. Phys. Chem. Interactions and Double Proton-Transfer Reactions at Both Gates of
A 2010, 114, 5618−5629. Cyclodextrins. Chem. Phys. Lett. 1998, 296, 335−342.
(165) Dantus, M.; Rosker, M. J.; Zewail, A. H. Real-Time (186) Alarcos, N.; Cohen, B.; Douhal, A. A Slowing Down of Proton
Femtosecond Probing of ‘‘Transition States’’in Chemical Reactions. Motion from Hpts to Water Adsorbed on the MCM-41 Surface. Phys.
J. Chem. Phys. 1987, 87, 2395−2397. Chem. Chem. Phys. 2016, 18, 2658−2671.
(166) Ernsting, N.; Kovalenko, S.; Senyushkina, T.; Saam, J.; (187) Smith, K. K.; Kaufmann, K. J.; Huppert, D.; Gutman, M.
Farztdinov, V. Wave-Packet-Assisted Decomposition of Femtosecond Picosecond Proton Ejection: An Ultrafast Ph Jump. Chem. Phys. Lett.
Transient Ultraviolet−Visible Absorption Spectra: Application to 1979, 64, 522−527.
Excited-State Intramolecular Proton Transfer in Solution. J. Phys. (188) Soroka, H. P.; Simkovitch, R.; Kosloff, A.; Shomer, S.; Pevzner,
Chem. A 2001, 105, 3443−3453. A.; Tzang, O.; Tirosh, R.; Patolsky, F.; Huppert, D. Excited-State

BR DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Proton Transfer and Proton Diffusion near Hydrophilic Surfaces. J. (208) Schmidhammer, U.; De Waele, V.; Mintova, S.; Riedle, E.;
Phys. Chem. C 2013, 117, 25786−25797. Bein, T. Femtochemistry of Guest Molecules Hosted in Colloidal
(189) Roy, D.; Karmakar, R.; Mondal, S. K.; Sahu, K.; Bhattacharyya, Zeolites. Adv. Funct. Mater. 2005, 15, 1973−1978.
K. Excited State Proton Transfer from Pyranine to Acetate in a CTAB (209) Cohen, B.; Wang, S.; Organero, J. A.; Campo, L. F.; Sanchez,
Micelle. Chem. Phys. Lett. 2004, 399, 147−151. F.; Douhal, A. Femtosecond Fluorescence Dynamics of a Proton-
(190) Mondal, S. K.; Sahu, K.; Sen, P.; Roy, D.; Ghosh, S.; Transfer Dye Interacting with Silica-Based Nanomaterials. J. Phys.
Bhattacharyya, K. Excited State Proton Transfer of Pyranine in a γ- Chem. C 2010, 114, 6281−6289.
Cyclodextrin Cavity. Chem. Phys. Lett. 2005, 412, 228−234. (210) Cohen, B.; Sanchez, F.; Douhal, A. Mapping the Distribution
(191) Cohen, B.; Martin Á lvarez, C.; Alarcos Carmona, N.; of an Individual Chromophore Interacting with Silica-Based Nano-
Organero, J. A.; Douhal, A. Proton-Transfer Reaction Dynamics materials. J. Am. Chem. Soc. 2010, 132, 5507−5514.
within the Human Serum Albumin Protein. J. Phys. Chem. B 2011, 115, (211) Grando, S. R.; Pessoa, C. M.; Gallas, M. R.; Costa, T. M. H.;
7637−7647. Rodembusch, F. S.; Benvenutti, E. V. Modulation of the ESIPT
(192) Hutter, T.; Presiado, I.; Ruschin, S.; Huppert, D. Protic Emission of Benzothiazole Type Dye Incorporated in Silica-Based
Properties of Water Confined in the Pores of Oxidized Porous Silicon Hybrid Materials. Langmuir 2009, 25, 13219−13223.
Studied by Excited-State Proton Transfer from a Photoacid. J. Phys. (212) Mintova, S.; De Waele, V.; Holzl, M.; Schmidhammer, U.;
Chem. C 2010, 114, 2341−2348. Mihailova, B.; Riedle, E.; Bein, T. Photochemistry of 2-(2′-
(193) Kim, T.-G.; Lee, S.-I.; Jang, D.-J.; Kim, Y. Unusually Large Hydroxyphenyl)Benzothiazole Encapsulated in Nanosized Zeolites. J.
Tunneling Effect on the Proton Transfer of Aqueous 7-Hydroxyquino- Phys. Chem. A 2004, 108, 10640−10648.
line. J. Phys. Chem. 1995, 99, 12698−12700. (213) Alarcos, N.; Cohen, B.; Douhal, A. Photodynamics of a Proton-
(194) Lee, S. Y.; Jang, D.-J. Proton Transfers of Aqueous 7- Transfer Dye in Solutions and Confined within NaX and NaY Zeolites.
Hydroxyquinoline in the First Excited Singlet, Lowest Triplet, and J. Phys. Chem. C 2014, 118, 19431−19443.
Ground States. J. Phys. Chem. 1995, 99, 7537−7541. (214) Grando, S. R.; da Silveira Santos, F.; Gallas, M. R.; Costa, T. M.
(195) Liu, Y.-H.; Mehata, M. S.; Liu, J.-Y. Excited-State Proton H.; Benvenutti, E. V.; Rodembusch, F. S. Photophysics of Amino-
Transfer Via Hydrogen-Bonded Acetic Acid (AcOH) Wire for 6- benzazole Dyes in Silica-Based Hybrid Materials. J. Sol-Gel Sci. Technol.
Hydroxyquinoline. J. Phys. Chem. A 2011, 115, 19−24. 2012, 63, 235−241.
(196) Park, S.-Y.; Kim, H.-B.; Yoo, B. K.; Jang, D.-J. Direct (215) Rodembusch, F. S.; Leusin, F. P.; Campo, L. F.; Stefani, V.
Observation of Conformation-Dependent Pathways in the Excited- Excited State Intramolecular Proton Transfer in Amino 2-(2′-
State Proton Transfer of 7-Hydroxyquinoline in Bulk Alcohols. J. Phys. Hydroxyphenyl)Benzazole Derivatives: Effects of the Solvent and
Chem. B 2012, 116, 14153−14158. the Amino Group Position. J. Lumin. 2007, 126, 728−734.
(197) Park, S.-Y.; Jang, D.-J. Excited-State Hydrogen Relay Along a (216) Santra, S.; Krishnamoorthy, G.; Dogra, S. K. Excited State
Blended-Alcohol Chain as a Model System of a Proton Wire: Intramolecular Proton Transfer in 2-(2′-Benzamidophenyl)-
Deuterium Effect on the Reaction Dynamics. Phys. Chem. Chem. Phys. Benzimidazole: Effect of Solvents. Chem. Phys. Lett. 1999, 311, 55−61.
2012, 14, 8885−8891. (217) Arthen-Engeland, T.; Bultmann, T.; Ernsting, N. P.; Rodriguez,
(198) Bach, A.; Tanner, C.; Manca, C.; Frey, H.-M.; Leutwyler, S. M. A.; Thiel, W. Singlet Excited-State Intramolecular Proton Tranfer
Ground- and Excited State Proton Transfer and Tautomerization in 7- in 2-(2t′-Hydroxyphenyl) Benzoxazole: Spectroscopy at Low Temper-
Hydroxyquinoline·(NH3)n Clusters: Spectroscopic and Time Resolved atures, Femtosecond Transient Absorption, and MNDO Calculations.
Investigations. J. Chem. Phys. 2003, 119, 5933−5942. Chem. Phys. 1992, 163, 43−53.
(199) Park, S.-Y.; Yu, H.; Park, J.; Jang, D.-J. Excited-State (218) Mordzinski, A.; Grellmann, K. H. Excited-State Proton-
Prototropic Equilibrium Dynamics of 6-Hydroxyquinoline Encapsu- Transfer Reactions in 2-(2′-Hydroxyphenyl)Benzoxazole. Role of
lated in Microporous Catalytic Faujasite Zeolites. Chem. - Eur. J. 2010, Triplet States. J. Phys. Chem. 1986, 90, 5503−5506.
16, 12609−12615. (219) Krishnamurthy, M.; Dogra, S. K. Proton Transfer of 2-(2′-
(200) Gil, M.; Organero, J. A.; Navarro, M. T.; Corma, A.; Douhal, A. Hydroxyphenyl)Benzothiazole in the Excited Single State. J. Photo-
Competitive Ultrafast Electron and Proton Transfer Reactions within chem. 1986, 32, 235−242.
Titania and Silica Mesoporous Materials. J. Phys. Chem. C 2012, 116, (220) Yang, W.; Chen, X. Dual Fluorescence of Excited State Intra-
15385−15395. Molecular Proton Transfer of HBFO: Mechanistic Understanding,
(201) Kim, T. G.; Kim, Y.; Jang, D.-J. Catalytic Roles of Water Substituent and Solvent Effects. Phys. Chem. Chem. Phys. 2014, 16,
Protropic Species in the Tautomerization of Excited 6-Hydroxyquino- 4242−4250.
line: Migration of Hydrated Proton Clusters. J. Phys. Chem. A 2001, (221) Alarcos, N.; Gutierrez, M.; Liras, M.; Sanchez, F.; Douhal, A.
105, 4328−4332. From Intra- to Inter-Molecular Hydrogen Bonds with the Surround-
(202) Lee, Y.-S.; Kwon, O.-H.; Park, H. J.; Franz, J.; Jang, D.-J. ings: Steady-State and Time-Resolved Behaviours. Photochem. Photo-
Excited-State Proton Transfer and Geminate Recombination in the biol. Sci. 2015, 14, 1306−1318.
Molecular Cage of β-Cyclodextrin. J. Photochem. Photobiol., A 2008, (222) Mohammed, O. F.; Luber, S.; Batista, V. S.; Nibbering, E. T. J.
194, 105−109. Ultrafast Branching of Reaction Pathways in 2-(2′-Hydroxyphenyl)-
(203) Poizat, O.; Bardez, E.; Buntinx, G.; Alain, V. Picosecond Benzothiazole in Polar Acetonitrile Solution. J. Phys. Chem. A 2011,
Dynamics of the Photoexcited 6-Methoxyquinoline and 6-Hydrox- 115, 7550−7558.
yquinoline Molecules in Solution. J. Phys. Chem. A 2004, 108, 1873− (223) Abou-Zied, O. K. The Role of Water in Solvating the
1880. Hydrogen-Bonding Center of 2-(2′-Hydroxyphenyl)Benzoxazole.
(204) Park, H. J.; Kwon, O.-H.; Ah, C. S.; Jang, D.-J. Excited-State Chem. Phys. 2007, 337, 1−10.
Tautomerization Dynamics of 7-Hydroxyquinoline in β-Cyclodextrin. (224) Wang, H.; Zhang, H.; Abou-Zied, O. K.; Yu, C.; Romesberg, F.
J. Phys. Chem. B 2005, 109, 3938−3943. E.; Glasbeek, M. Femtosecond Fluorescence Upconversion Studies of
(205) Park, S.-Y.; Kwon, O.-H.; Kim, T. G.; Jang, D.-J. Ground-State Excited-State Proton-Transfer Dynamics in 2-(2′-Hydroxyphenyl)-
Proton Transfer of 7-Hydroxyquinoline Confined in Biologically Benzoxazole (HBO) in Liquid Solution and DNA. Chem. Phys. Lett.
Relevant Water Nanopools. J. Phys. Chem. C 2009, 113, 16110−16115. 2003, 367, 599−608.
(206) Simkovitch, R.; Shomer, S.; Gepshtein, R.; Huppert, D. How (225) Elsaesser, T.; Schmetzer, B. Excited-State Proton Transfer in 2-
Fast Can a Proton-Transfer Reaction Be Beyond the Solvent-Control (2′-Hydroxyphenyl)Benzothiazole Formation of the Anion in Polar
Limit? J. Phys. Chem. B 2015, 119, 2253−2262. Solvents. Chem. Phys. Lett. 1987, 140, 293−299.
(207) Cohen, B.; Segal, J.; Huppert, D. Proton Transfer from (226) Martin, C.; Cohen, B.; Navarro, M. T.; Corma, A.; Douhal, A.
Photoacid to Solvent. J. Phys. Chem. A 2002, 106, 7462−7467. Unraveling the Ultrafast Behavior of Nile Red Interacting with

BS DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Aluminum and Titanium Co-Doped MCM41 Materials. Phys. Chem. (244) Viswanathan, K.; Natarajan, P. Photophysical Properties of
Chem. Phys. 2016, 18, 2152−2163. Thionine and Phenosafranine Dyes Covalently Bound to Macro-
(227) Martín, C.; Piatkowski, P.; Cohen, B.; Gil, M.; Navarro, M. T.; molecules. J. Photochem. Photobiol., A 1996, 95, 245−253.
Corma, A.; Douhal, A. Ultrafast Dynamics of Nile Red Interacting with (245) Ferreira, L. F. V.; Ferreira, M. R. V.; Oliveira, A. S.; Branco, T.
Metal Doped Mesoporous Materials. J. Phys. Chem. C 2015, 119, J. F.; Prata, J. V.; Moreira, J. C. Diffuse Reflectance Studies of β-
13283−13296. Phenylpropiophenone and Benzophenone Inclusion Complexes with
(228) Konovalova, T. A.; Gao, Y.; Schad, R.; Kispert, L. D.; Saylor, C. Calix[4], [6] and [8]Arenes. Phys. Chem. Chem. Phys. 2002, 4, 204−
A.; Brunel, L.-C. Photooxidation of Carotenoids in Mesoporous 210.
MCM-41, Ni- MCM-41 and Al- MCM-41 Molecular Sieves. J. Phys. (246) Vieira Ferreira, L. F.; Netto-Ferreira, J. C.; Costa, S. M. B.
Chem. B 2001, 105, 7459−7464. Time-Resolved Absorption and Emission Spectra of Triplet State β-
(229) Kumar, K. S.; Selvaraju, C.; Malar, E. J. P.; Natarajan, P. Phenylpropiophenone Adsorbed on Silicalite. Spectrochim. Acta, Part A
Existence of a New Emitting Singlet State of Proflavine: Femtosecond 1995, 51, 1385−1388.
Dynamics of the Excited State Processes and Quantum Chemical (247) Ferreira, L. F. V.; Ferreira, M. R. V.; Da Silva, J. P.; Ferreira
Studies in Different Solvents. J. Phys. Chem. A 2012, 116, 37−45. Machado, I.; Oliveira, A. S.; Prata, J. V. Novel Laser-Induced
(230) Ananthanarayanan, K.; Natarajan, P. Fabrication and Photo- Luminescence Resulting from Benzophenone/o-Propylated p-Tert-
physical Studies of Phenosafranine and Proflavine Dyes Encapsulated Butylcalix[4]Arene Complexes. A Diffuse Reflectance Study. Photo-
in Mesoporous MCM-41 Along with Titanium Dioxide Nanoparticles. chem. Photobiol. Sci. 2003, 2, 1002−1010.
Microporous Mesoporous Mater. 2009, 124, 179−189. (248) Vieira Ferreira, L. F.; Vieira Ferreira, M. R.; Oliveira, A. S.;
(231) Ananthanarayanan, K.; Selvaraju, C.; Natarajan, P. Novel Moreira, J. C. Potentialities of Diffuse Reflectance Laser-Induced
Excited State Proton Transfer Reaction Observed for Proflavine Techniques in Solid Phase: A Comparative Study of Benzophenone
Encapsulated in the Channels of Modified MCM-41. Microporous Inclusion within p-tert-Butylcalixarenes, Silicalite and Microcrystalline
Mesoporous Mater. 2007, 99, 319−327. Cellulose. J. Photochem. Photobiol., A 2002, 153, 11−18.
(232) Senthilkumar, K.; Chandra Mohan, S.; Easwaramoorthi, S.; (249) Ghosh, R.; Palit, D. K. Dynamics of Solvent Controlled Excited
Jothivenkatachalam, K.; Natarajan, P. Photoprocesses of Molecules State Intramolecular Proton Transfer Coupled Charge Transfer
Encapsulated in Porous Solids XI: Excited State Dynamics of Reactions. Photochem. Photobiol. Sci. 2013, 12, 987−995.
Proflavine and Photosensitization of TiO2 in Nanoporous Materials. (250) Kim, S.; Seo, J.; Park, S. Y. Torsion-Induced Fluorescence
Microporous Mesoporous Mater. 2014, 195, 124−130. Quenching in Excited-State Intramolecular Proton Transfer (ESIPT)
(233) Natarajan, P.; Duraimurugan, K.; Kumar, K. S. Photoprocesses Dyes. J. Photochem. Photobiol., A 2007, 191, 19−24.
of Coordination Compounds and Dyes in Solution and Nanoporous (251) Park, S.; Kwon, J. E.; Park, S. Y. Strategic Emission Color
Materials: Evolution from Milliseconds to Femtoseconds. J. Chem. Sci. Tuning of Highly Fluorescent Imidazole-Based Excited-State Intra-
molecular Proton Transfer Molecules. Phys. Chem. Chem. Phys. 2012,
2011, 123, 531−553.
14, 8878−8884.
(234) Senthilkumar, K.; Paul, P.; Selvaraju, C.; Natarajan, P.
(252) Sliwa, M.; Mouton, N.; Ruckebusch, C.; Aloïse, S.; Poizat, O.;
Preparation, Characterization, and Photophysical Study of Thiazine
Buntinx, G.; Métivier, R.; Nakatani, K.; Masuhara, H.; Asahi, T.
Dyes within the Nanotubes and Nanoavities of Silicate Host: Influence
Comparative Investigation of Ultrafast Photoinduced Processes in
of Titanium Dioxide Nanoparticle on the Protonation and Aggregation
Salicylidene-Aminopyridine in Solution and Solid State. J. Phys. Chem.
of Dyes. J. Phys. Chem. C 2010, 114, 7085−7094.
C 2009, 113, 11959−11968.
(235) Gigli, L.; Arletti, R.; Vitillo, J. G.; Alberto, G.; Martra, G.;
(253) Sun, W.; Li, S.; Hu, R.; Qian, Y.; Wang, S.; Yang, G.
Devaux, A.; Vezzalini, G. Thionine Dye Confined in Zeolite L: Understanding Solvent Effects on Luminescent Properties of a Triple
Synthesis Location and Optical Properties. J. Phys. Chem. C 2015, 119, Fluorescent ESIPT Compound and Application for White Light
16156−16165. Emission. J. Phys. Chem. A 2009, 113, 5888−5895.
(236) Da Silva, J. P.; Ferreira Machado, I.; Lourenço, J. P.; Vieira (254) Yushchenko, D. A.; Shvadchak, V. V.; Klymchenko, A. S.;
Ferreira, L. F. Photochemistry of Benzophenone Adsorbed on MCM- Duportail, G.; Pivovarenko, V. G.; Mély, Y. Modulation of Excited-
41 Surface. Microporous Mesoporous Mater. 2005, 84, 1−10. State Intramolecular Proton Transfer by Viscosity in Protic Media. J.
(237) Ferreira, L. F. V.; Machado, I. F.; Da Silva, J. P.; Branco, T. J. F. Phys. Chem. A 2007, 111, 10435−10438.
Surface Photochemistry: Benzophenone as a Probe for the Study of (255) Ikeda, T.; Tsutsumi, O. Optical Switching and Image Storage
Silica and Reversed-Phase Silica Surfaces. Photochem. Photobiol. Sci. by Means of Azobenzene Liquid-Crystal Films. Science 1995, 268,
2006, 5, 665−673. 1873−1875.
(238) Vieira Ferreira, L. F.; Costa, A. I.; Ferreira Machado, I.; Da (256) Liu, Z. F.; Hashimoto, K.; Fujishima, A. Photoelectrochemical
Silva, J. P. Surface Photochemistry: Benzophenone within Nano- Information Storage Using an Azobenzene Derivative. Nature 1990,
channels of H+ and Na+ ZSM-5 Zeolites. Microporous Mesoporous 347, 658−660.
Mater. 2009, 119, 82−90. (257) Conti, I.; Garavelli, M.; Orlandi, G. The Different Photo-
(239) Machado, I. F.; Vieira Ferreira, L. F.; Branco, T. J. F.; isomerization Efficiency of Azobenzene in the Lowest nπ* and ππ*
Fernandes, A.; Ribeiro, F. Surface Photochemistry: Ketones Included Singlets: The Role of a Phantom State. J. Am. Chem. Soc. 2008, 130,
within a Channel Type Solid Support, the Aluminophosphate AlPO4- 5216−5230.
5. J. Mol. Struct. 2007, 831, 1−9. (258) Lu, Y.-C.; Diau, E. W.-G.; Rau, H. Femtosecond Fluorescence
(240) Vieira Ferreira, L. F.; Duarte, P.; Ferreira, D. P.; Ferreira Dynamics of Rotation-Restricted Azobenzenophanes: New Evidence
Machado, I.; Oliveira, A. S.; Prukała, D.; Sikorski, M. Surface on the Mechanism of Trans → Cis Photoisomerization of
Photochemistry: p-Hydroxystilbazol within Nanochannels of Na+ Azobenzene. J. Phys. Chem. A 2005, 109, 2090−2099.
and H+ ZSM-5 Zeolites. Microporous Mesoporous Mater. 2012, 151, (259) Joshi, H.; Kamounah, F. S.; Gooijer, C.; van der Zwan, G.;
317−324. Antonov, L. Excited State Intramolecular Proton Transfer in Some
(241) Pileni, M. P.; Graetzel, M. Light-Induced Redox Reactions of Tautomeric Azo Dyes and Schiff Bases Containing an Intramolecular
Proflavine in Aqueous and Micellar Solution. J. Phys. Chem. 1980, 84, Hydrogen Bond. J. Photochem. Photobiol., A 2002, 152, 183−191.
2402−2406. (260) Gil, M.; Wang, S.; Organero, J. A.; Teruel, L.; Garcia, H.;
(242) Scaiano, J. C.; García, H. Intrazeolite Photochemistry: Toward Douhal, A. Femtosecond Dynamics within Nanotubes and Nano-
Supramolecular Control of Molecular Photochemistry. Acc. Chem. Res. cavities of Mesoporous and Zeolite Materials. J. Phys. Chem. C 2009,
1999, 32, 783−793. 113, 11614−11622.
(243) Turro, N. J. Photochemistry of Organic Molecules in (261) Gil, M.; Organero, J.; Peris, E.; Garcia, H.; Douhal, A.
Microscopic Reactors. Pure Appl. Chem. 1986, 58, 1219−1228. Confinement Effect of Nanocages and Nanotubes of Mesoporous

BT DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Materials on the Keto Forms Photodynamics of Sudan I. Chem. Phys. (281) Blumberger, J. Recent Advances in the Theory and Molecular
Lett. 2009, 474, 325−330. Simulation of Biological Electron Transfer Reactions. Chem. Rev. 2015,
(262) Douhal, A.; Sanz, M.; Tormo, L. Femtochemistry of Orange II 115, 11191−11238.
in Solution and in Chemical and Biological Nanocavities. Proc. Natl. (282) Marcus, R. A. On the Theory of Oxidation-Reduction
Acad. Sci. U. S. A. 2005, 102, 18807−18812. Reactions Involving Electron Transfer. I. J. Chem. Phys. 1956, 24,
(263) Gil, M.; Ziółek, M.; Organero, J. A.; Douhal, A. Confined Fast 966−978.
and Ultrafast Dynamics of a Photochromic Proton-Transfer Dye (283) Marcus, R. Chemical and Electrochemical Electron-Transfer
within a Zeolite Nanocage. J. Phys. Chem. C 2010, 114, 9554−9562. Theory. Annu. Rev. Phys. Chem. 1964, 15, 155−196.
(264) Alarcos, N.; Organero, J. A.; Sánchez, F.; Douhal, A. Exploring (284) Kavarnos, G. J. Fundamental Concepts of Photoinduced
the Photobehavior of Nanocaged Monomers and H- and J-Aggregates Electron Transfer. In Photoinduced Electron Transfer I; Springer: Berlin,
of a Proton-Transfer Dye within NaX and NaY Zeolites. J. Phys. Chem. Heidelberg, 1990; Vol 156, pp 21−58.
C 2014, 118, 8217−8226. (285) Barbara, P. F.; Jarzeba, W. Ultrafast Photochemical Intra-
(265) Alarcos, N.; Sánchez, F.; Douhal, A. Spectroscopy and molecular Charge and Excited State Solvation. Adv. Photochem. 1990,
Relaxation Dynamics of Salicylideneaniline Derivative Aggregates 15, 1−68.
Encapsulated in MCM41 and SBA15 Pores. Microporous Mesoporous (286) Marcus, R. A. Electron Transfer Reactions in Chemistry:
Theory and Experiment (Nobel Lecture). Angew. Chem., Int. Ed. Engl.
Mater. 2016, 226, 34−43.
1993, 32, 1111−1121.
(266) Alarcos, N.; Sánchez, F.; Douhal, A. Interrogating Ultrafast
(287) Chipem, F. A. S.; Mishra, A.; Krishnamoorthy, G. The Role of
Dynamics of a Salicylideneaniline Derivative within Faujasite Zeolites.
Hydrogen Bonding in Excited State Intramolecular Charge Transfer.
Chem. Phys. Lett. 2017, 683, 145−153. Phys. Chem. Chem. Phys. 2012, 14, 8775−8790.
(267) Alarcos, N.; Sánchez, F.; Douhal, A. Confinement Effect on (288) Karunakaran, V.; Das, S. Direct Observation of Cascade of
Ultrafast Events of a Salicylideneaniline Derivative within Mesoporous Photoinduced Ultrafast Intramolecular Charge Transfer Dynamics in
Materials. Microporous Mesoporous Mater. 2017, 248, 54−61. Diphenyl Acetylene Derivatives: Via Solvation and Intramolecular
(268) Mondal, S.; Basu, S.; Mandal, D. Ground- and Excited-State Relaxation. J. Phys. Chem. B 2016, 120, 7016−7023.
Proton-Transfer Reaction of 3-Hydroxyflavone in Aqueous Micelles. (289) Stoltzfus, D. M.; Donaghey, J. E.; Armin, A.; Shaw, P. E.; Burn,
Chem. Phys. Lett. 2009, 479, 218−223. P. L.; Meredith, P. Charge Generation Pathways in Organic Solar
(269) Tormo, L.; Douhal, A. Caging Anionic Structure of a Proton Cells: Assessing the Contribution from the Electron Acceptor. Chem.
Transfer Dye in a Hydrophobic Nanocavity with a Cooperative H- Rev. 2016, 116, 12920−12955.
Bonding. J. Photochem. Photobiol., A 2005, 173, 358−364. (290) Wasielewski, M. R. Photoinduced Electron Transfer in
(270) Banerjee, A.; Sengupta, P. K. Encapsulation of 3-Hydroxy- Supramolecular Systems for Artificial Photosynthesis. Chem. Rev.
flavone and Fisetin in β-Cyclodextrins: Excited State Proton Transfer 1992, 92, 435−461.
Fluorescence and Molecular Mechanics Studies. Chem. Phys. Lett. (291) Wiedbrauk, S.; Maerz, B.; Samoylova, E.; Reiner, A.; Trommer,
2006, 424, 379−386. F.; Mayer, P.; Zinth, W.; Dube, H. Twisted Hemithioindigo
(271) Moissette, A.; Hureau, M.; Kokaislova, A.; Le Person, A.; Photoswitches: Solvent Polarity Determines the Type of Light-
Cornard, J. P.; De Waele, I.; Batonneau-Gener, I. Spectroscopic Induced Rotations. J. Am. Chem. Soc. 2016, 138, 12219−12227.
Evidence of 3-Hydroxyflavone Sorption within MFI Type Zeolites: (292) Ya Freidzon, A.; Safonov, A. A.; Bagaturyants, A. A.; Alfimov,
ESIPT and Metal Complexation. Phys. Chem. Chem. Phys. 2015, 17, M. V. Solvatofluorochromism and Twisted Intramolecular Charge-
26207−26219. Transfer State of the Nile Red Dye. Int. J. Quantum Chem. 2012, 112,
(272) Doussineau, T.; Smaïhi, M.; Balme, S.; Janot, J.-M. Fluorescent 3059−3067.
Hydroxyflavone−Zeolite Nanoparticles: Ship-in-a-Bottle Synthesis and (293) Sarkar, N.; Das, K.; Nath, D. N.; Bhattacharyya, K. Twisted
Photophysical Properties. ChemPhysChem 2006, 7, 583−589. Charge Transfer Process of Nile Red in Homogeneous Solution and in
(273) Ziółek, M.; Sobczak, I. Photochromism and Hydrolysis of Faujasite Zeolite. Langmuir 1994, 10, 326−329.
Aromatic Schiff Base N,N′-Bis(Salicylidene)-p-Phenylenediamine (294) Sackett, D. L.; Wolff, J. Nile Red as a Polarity-Sensitive
(BSP) Studied in Heterogeneous Environments. J. Inclusion Phenom. Fluorescent Probe of Hydrophobic Protein Surfaces. Anal. Biochem.
Mol. Recognit. Chem. 2009, 63, 211−218. 1987, 167, 228−234.
(274) Strandjord, A. J. G.; Barbara, P. F. The Proton-Transfer (295) Dutta, A. K.; Kamada, K.; Ohta, K. Spectroscopic Studies of
Kinetics of 3-Hydroxyflavone: Solvent Effects. J. Phys. Chem. 1985, 89, Nile Red in Organic Solvents and Polymers. J. Photochem. Photobiol., A
2355−2361. 1996, 93, 57−64.
(275) McMorrow, D.; Kasha, M. Intramolecular Excited-State Proton (296) Tran-Thi, T. H.; Gustavsson, T.; Prayer, C.; Pommeret, S.;
Hynes, J. T. Primary Ultrafast Events Preceding the Photoinduced
Transfer in 3-Hydroxyflavone. Hydrogen-Bonding Solvent Perturba-
Proton Transfer from Pyranine to Water. Chem. Phys. Lett. 2000, 329,
tions. J. Phys. Chem. 1984, 88, 2235−2243.
421−430.
(276) Carturan, S.; Quaranta, A.; Maggioni, G.; Vomiero, A.;
(297) Tran-Thi, T. H.; Prayer, C.; Millié, P.; Uznanski, P.; Hynes, J.
Ceccato, R.; Mea, G. D. Optical Study of the Matrix Effect on the
T. Substituent and Solvent Effects on the Nature of the Transitions of
ESIPT Mechanism of 3-HF Doped Sol-Gel Glass. J. Sol-Gel Sci. Pyrenol and Pyranine. Identification of an Intermediate in the Excited-
Technol. 2003, 26, 931−935. State Proton-Transfer Reaction. J. Phys. Chem. A 2002, 106, 2244−
(277) Protti, S.; Raulin, K.; Cristini, O.; Kinowski, C.; Turrell, S.; 2255.
Mezzetti, A. Wavelength Shifting Systems Based on Flavonols and (298) Mondal, T.; Das, A. K.; Sasmal, D. K.; Bhattacharyya, K.
Their Metal Complexes Encapsulated by Post-Doping in Porous SiO2 Excited State Proton Transfer in Ionic Liquid Mixed Micelles. J. Phys.
Xerogel Matrices. J. Mol. Struct. 2011, 993, 485−490. Chem. B 2010, 114, 13136−13142.
(278) Guharay, J.; Chaudhuri, R.; Chakrabarti, A.; Sengupta, P. K. (299) Ghosh, S.; Dey, S.; Mandal, U.; Adhikari, A.; Mondal, S. K.;
Excited State Proton Transfer Fluorescence of 3-Hydroxyflavone in Bhattacharyya, K. Ultrafast Proton Transfer of Pyranine in a
Model Membranes. Spectrochim. Acta, Part A 1997, 53, 457−462. Supramolecular Assembly: PEO−PPO−PEO Triblock Copolymer
(279) Shyamala, T.; Mishra, A. K. Ground- and Excited-State Proton and CTAC. J. Phys. Chem. B 2007, 111, 13504−13510.
Transfer Reaction of 3-Hydroxyflavone in Dimyristoylphosphatidyl- (300) Pal, S. K.; Peon, J.; Bagchi, B.; Zewail, A. H. Biological Water:
choline Liposome Membrane. Photochem. Photobiol. 2004, 80, 309− Femtosecond Dynamics of Macromolecular Hydration. J. Phys. Chem.
315. B 2002, 106, 12376−12395.
(280) Marcus, R. A.; Sutin, N. Electron Transfers in Chemistry and (301) Fayer, M. D.; Levinger, N. E. Analysis of Water in Confined
Biology. Biochim. Biophys. Acta, Rev. Bioenerg. 1985, 811, 265−322. Geometries and at Interfaces. Annu. Rev. Anal. Chem. 2010, 3, 89−107.

BU DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(302) Thompson, W. H. A Monte Carlo Study of Spectroscopy in Sum Frequency Generation Spectroscopy of the Alumina(0001)/
Nanoconfined Solvents. J. Chem. Phys. 2002, 117, 6618−6628. Water Interface. J. Phys. Chem. C 2017, 121, 5168−5177.
(303) Sansom, M. S.; Kerr, I. D.; Breed, J.; Sankararamakrishnan, R. (323) Ziółek, M.; Martín, C.; Navarro, M. T.; Garcia, H.; Douhal, A.
Water in Channel-Like Cavities: Structure and Dynamics. Biophys. J. Confined Photodynamics of an Organic Dye for Solar Cells
1996, 70, 693−702. Encapsulated in Titanium-Doped Mesoporous Molecular Materials.
(304) Mondal, S. K.; Sahu, K.; Ghosh, S.; Sen, P.; Bhattacharyya, K. J. Phys. Chem. C 2011, 115, 8858−8867.
Excited-State Proton Transfer from Pyranine to Acetate in γ- (324) Hwang, S.; Lee, J. H.; Park, C.; Lee, H.; Kim, C.; Park, C.; Lee,
Cyclodextrin and Hydroxypropyl γ-Cyclodextrin. J. Phys. Chem. A M.-H.; Lee, W.; Park, J.; Kim, K.; et al. A Highly Efficient Organic
2006, 110, 13646−13652. Sensitizer for Dye-Sensitized Solar Cells. Chem. Commun. 2007,
(305) Stratt, R. M.; Maroncelli, M. Nonreactive Dynamics in 4887−4889.
Solution: The Emerging Molecular View of Solvation Dynamics and (325) Tian, H.; Yang, X.; Chen, R.; Zhang, R.; Hagfeldt, A.; Sun, L.
Vibrational Relaxation. J. Phys. Chem. 1996, 100, 12981−12996. Effect of Different Dye Baths and Dye-Structures on the Performance
(306) Pal, S. K.; Sukul, D.; Mandal, D.; Sen, S.; Bhattacharyya, K. of Dye-Sensitized Solar Cells Based on Triphenylamine Dyes. J. Phys.
Solvation Dynamics of Coumarin 480 in Sol−Gel Matrix. J. Phys. Chem. C 2008, 112, 11023−11033.
Chem. B 2000, 104, 2613−2616. (326) Ziółek, M.; Martín, C.; Cohen, B.; Garcia, H.; Douhal, A.
(307) Baumann, R.; Ferrante, C.; Deeg, F. W.; Bräuchle, C. Solvation Virtues and Vices of an Organic Dye and Ti-Doped MCM-41 Based
Dynamics of Nile Blue in Ethanol Confined in Porous Sol−Gel Dye-Sensitized Solar Cells. J. Phys. Chem. C 2011, 115, 23642−23650.
Glasses. J. Chem. Phys. 2001, 114, 5781−5791. (327) Martín, C.; Ziółek, M.; Marchena, M.; Douhal, A. Interfacial
(308) Baumann, R.; Ferrante, C.; Kneuper, E.; Deeg, F.-W.; Bräuchle, Electron Transfer Dynamics in a Solar Cell Organic Dye Anchored to
C. Influence of Confinement on the Solvation and Rotational Semiconductor Particle and Aluminum-Doped Mesoporous Materials.
Dynamics of Coumarin 153 in Ethanol. J. Phys. Chem. A 2003, 107, J. Phys. Chem. C 2011, 115, 23183−23191.
2422−2430. (328) di Nunzio, M. R.; Cohen, B.; Pandey, S.; Hayse, S.; Piani, G.;
(309) Nandi, N.; Bagchi, B. Ultrafast Solvation Dynamics of an Ion in Douhal, A. Spectroscopy and Dynamics of Yd2-O-C8 in Solution and
the γ-Cyclodextrin Cavity: The Role of Restricted Environment. J. Interacting with Alumina Nanoparticles Electrode. J. Phys. Chem. C
Phys. Chem. 1996, 100, 13914−13919. 2014, 118, 11365−11376.
(310) Vajda, S.; Jimenez, R.; Rosenthal, S. J.; Fidler, V.; Fleming, G. (329) Hsu, H.-Y.; Chiang, H.-C.; Hu, J.-Y.; Awasthi, K.; Mai, C.-L.;
R.; Castner, E. W. Femtosecond to Nanosecond Solvation Dynamics Yeh, C.-Y.; Ohta, N.; Diau, E. W.-G. Field-Induced Fluorescence
in Pure Water and inside the γ-Cyclodextrin Cavity. J. Chem. Soc., Quenching and Enhancement of Porphyrin Sensitizers on TiO2 Films
Faraday Trans. 1995, 91, 867−873. and in PMMA Films. J. Phys. Chem. C 2013, 117, 24761−24766.
(311) Sarkar, N.; Datta, A.; Das, S.; Bhattacharyya, K. Solvation (330) Palomares, E.; Clifford, J. N.; Haque, S. A.; Lutz, T.; Durrant, J.
Dynamics of Coumarin 480 in Micelles. J. Phys. Chem. 1996, 100, R. Control of Charge Recombination Dynamics in Dye Sensitized
Solar Cells by the Use of Conformally Deposited Metal Oxide
15483−15486.
Blocking Layers. J. Am. Chem. Soc. 2003, 125, 475−482.
(312) Yamaguchi, A.; Amino, Y.; Shima, K.; Suzuki, S.; Yamashita, T.;
(331) Luo, L.; Lin, C.-J.; Tsai, C.-Y.; Wu, H.-P.; Li, L.-L.; Lo, C.-F.;
Teramae, N. Local Environments of Coumarin Dyes within
Lin, C.-Y.; Diau, E. W.-G. Effects of Aggregation and Electron
Mesostructured Silica−Surfactant Nanocomposites. J. Phys. Chem. B
Injection on Photovoltaic Performance of Porphyrin-Based Solar Cells
2006, 110, 3910−3916.
with Oligo(Phenylethynyl) Links inside TiO2 and Al2O3 Nanotube
(313) Paul, A.; Sarkar, M.; Khara, D. C.; Kamijo, T.; Yamaguchi, A.;
Arrays. Phys. Chem. Chem. Phys. 2010, 12, 1064−1071.
Teramae, N.; Samanta, A. Solvation Dynamics of a Surfactant Probe in (332) Belogorokhov, I. A.; Ryabchikov, Y. V.; Tikhonov, E. V.;
Mesostructured Silica-Surfactant Nanocomposites. Chem. Phys. Lett. Pushkarev, V. E.; Breusova, M. O.; Tomilova, L. G.; Khokhlov, D. R.
2009, 469, 71−75. Photoluminescence in Semiconductor Structures Based on Butyl-
(314) Datta, A.; Mandal, D.; Pal, S. K.; Das, S.; Bhattacharyya, K. Substituted Erbium Phthalocyanine Complexes. Semiconductors 2008,
Solvation Dynamics in Organized Assemblies, 4-Aminophthalimide in 42, 321−324.
Micelles. J. Mol. Liq. 1998, 77, 121−129. (333) Wöhrle, D.; Meissner, D. Organic Solar Cells. Adv. Mater.
(315) Das, S.; Datta, A.; Bhattacharyya, K. Deuterium Isotope Effect 1991, 3, 129−138.
on 4-Aminophthalimide in Neat Water and Reverse Micelles. J. Phys. (334) Walzer, K.; Maennig, B.; Pfeiffer, M.; Leo, K. Highly Efficient
Chem. A 1997, 101, 3299−3304. Organic Devices Based on Electrically Doped Transport Layers. Chem.
(316) Kamijo, T.; Yamaguchi, A.; Suzuki, S.; Teramae, N.; Itoh, T.; Rev. 2007, 107, 1233−1271.
Ikeda, T. Solvation Dynamics of Coumarin 153 in Alcohols Confined (335) Zhou, R.; Josse, F.; Göpel, W.; Ö ztürk, Z. Z.; Bekaroğlu, Ö .
in Silica Nanochannels. J. Phys. Chem. A 2008, 112, 11535−11542. Phthalocyanines as Sensitive Materials for Chemical Sensors. Appl.
(317) Hazra, P.; Chakrabarty, D.; Chakraborty, A.; Sarkar, N. Organomet. Chem. 1996, 10, 557−577.
Intramolecular Charge Transfer and Solvation Dynamics of Nile Red (336) Singh, S.; Tripathi, S. K.; Saini, G. S. S. Optical and Infrared
in the Nanocavity of Cyclodextrins. Chem. Phys. Lett. 2004, 388, 150− Spectroscopic Studies of Chemical Sensing by Copper Phthalocyanine
157. Thin Films. Mater. Chem. Phys. 2008, 112, 793−797.
(318) Datta, A.; Mandal, D.; Pal, S. K.; Bhattacharyya, K. (337) Wang, S.; Liu, Y.; Huang, X.; Yu, G.; Zhu, D. Phthalocyanine
Intramolecular Charge Transfer Processes in Confined Systems. Nile Monolayer-Modified Gold Substrates as Efficient Anodes for Organic
Red in Reverse Micelles. J. Phys. Chem. B 1997, 101, 10221−10225. Light-Emitting Diodes. J. Phys. Chem. B 2003, 107, 12639−12642.
(319) Shen, Y. R.; Ostroverkhov, V. Sum-Frequency Vibrational (338) Synak, A.; Gil, M.; Organero, J. A.; Sánchez, F.; Iglesias, M.;
Spectroscopy on Water Interfaces: Polar Orientation of Water Douhal, A. Fast to Ultrafast Dynamics of Palladium Phthalocyanine
Molecules at Interfaces. Chem. Rev. 2006, 106, 1140−1154. Covalently Bonded to MCM-41 Mesoporous Material. J. Phys. Chem.
(320) Eftekhari-Bafrooei, A.; Borguet, E. Effect of Surface Charge on C 2009, 113, 19199−19207.
the Vibrational Dynamics of Interfacial Water. J. Am. Chem. Soc. 2009, (339) Fukuzumi, S.; Itoh, A.; Suenobu, T.; Ohkubo, K. Formation of
131, 12034−12035. the Long-Lived Charge-Separated State of the 9-Mesityl-10-Methyl-
(321) Boulesbaa, A.; Borguet, E. Vibrational Dynamics of Interfacial acridinium Cation Incorporated into Mesoporous Aluminosilicate at
Water by Free Induction Decay Sum Frequency Generation (FID- High Temperatures. J. Phys. Chem. C 2014, 118, 24188−24196.
SFG) at the Al2O3(1120)/H2O Interface. J. Phys. Chem. Lett. 2014, 5, (340) Lobo, R. F.; Moissette, A.; Hureau, M.; Carré, S.; Vezin, H.;
528−533. Legrand, A. Electron Transfers Induced by t-Stilbene Sorption in
(322) Tuladhar, A.; Piontek, S. M.; Borguet, E. Insights on Interfacial Acidic Aluminum, Gallium, and Boron Beta (BEA) Zeolites. J. Phys.
Structure, Dynamics, and Proton Transfer from Ultrafast Vibrational Chem. C 2012, 116, 14480−14490.

BV DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(341) Hureau, M.; Moissette, A.; Legrand, A.; Luchez, F.; Sliwa, M.; Application Using a Probe Chromophore Molecule. Phys. Chem. Chem.
Bremard, C. Chemical Control of Photoinduced Charges under Phys. 2014, 16, 13145−13155.
Confinement in Zeolites. J. Phys. Chem. C 2012, 116, 9092−9105. (359) Iu, K.-K.; Liu, X.; Thomas, J. K. Photoionization of
(342) Fukuzumi, S.; Doi, K.; Itoh, A.; Suenobu, T.; Ohkubo, K.; Naphthalene, Trans-Stilbene and Anthracene Adsorbed on the
Yamada, Y.; Karlin, K. D. Formation of a Long-Lived Electron- External and Internal Surfaces of Zeolites. J. Photochem. Photobiol., A
Transfer State in Mesoporous Silica-Alumina Composites Enhances 1994, 79, 103−107.
Photocatalytic Oxygenation Reactivity. Proc. Natl. Acad. Sci. U. S. A. (360) Keirstead, A. E.; Schepp, N. P.; Cozens, F. L. Influence of the
2012, 109, 15572−15577. Alkali Metal Cation on the Distance of Electron Migration in Zeolite
(343) Luchez, F.; Carre, S.; Moissette, A.; Poizat, O. Sorption and Y: A Nanosecond Laser Photolysis Study. J. Phys. Chem. C 2007, 111,
Spontaneous Ionization of Phenothiazine within Channel Type 14247−14252.
Zeolites: Effect of the Confinement on the Electron Transfers. RSC (361) Nakato, T.; Watanabe, S.; Kamijo, Y.; Nono, Y. Photoinduced
Adv. 2011, 1, 341−350. Electron Transfer between Ruthenium-Bipyridyl Complex and
(344) Dutta, P. K.; Severance, M. Photoelectron Transfer in Zeolite Methylviologen in Suspensions of Smectite Clays. J. Phys. Chem. C
Cages and Its Relevance to Solar Energy Conversion. J. Phys. Chem. 2012, 116, 8562−8570.
Lett. 2011, 2, 467−476. (362) Corma, A.; Fornes, V.; Galletero, M. S.; Garcia, H.; Scaiano, J.
(345) Moissette, A.; Lobo, R. F.; Vezin, H.; Al-Majnouni, K. A.; C. Evidence for through-Framework Electron Transfer in Intrazeolite
Brémard, C. Long Lived Charge Separated States Induced by Trans- Photochemistry. Case of Ru(Bpy)32+ and Methylviologen in Novel
Stilbene Incorporation in the Pores of Brønsted Acidic HZSM-5 Delaminated ITQ-2 Zeolite. Chem. Commun. 2002, 334−335.
Zeolites: Effect of Gallium on the Spontaneous Ionization Process. J. (363) Zhang, H.; Rajesh, C. S.; Dutta, P. K. Ruthenium Polypyridyl
Phys. Chem. C 2010, 114, 10280−10290. Complexes Containing a Conjugated Ligand LDQ(LDQ = 1-[4-(4′-
(346) Moissette, A.; Belhadj, F.; Brémard, C.; Vezin, H. Kinetics and methyl)-2,2′-bipyridyl]-2-[4-(4′-N,N′-tetramethylene-2,2′-
Characterization of Photoinduced Long-Lived Electron-Hole Pair of p- bipyridinum)]ethene): Synthesis, Characterization, and Photoinduced
Terphenyl Occluded in ZSM-5 Zeolites. Effects of Aluminium Electron Transfer at Solution−Zeolite Interfaces. J. Phys. Chem. C
Content and Extraframework Cation. Phys. Chem. Chem. Phys. 2009, 2009, 113, 4623−4633.
11, 11022−11032. (364) Kumar, K. S.; Sudha, T. S.; Natarajan, P. Photoprocess of
(347) Moissette, A.; Brémard, C.; Hureau, M.; Vezin, H. Slow Molecules Encapsulated in Porous Solids X: Photosensitization of
Interfacial Electron Hole Transfer of a Trans-Stilbene Radical Cation Zeolite-Y Encapsulated Tris(2,2′-Bipyridine- Nickel-(II)Ion by
Photoinduced in a Channel of Nonacidic Aluminum Rich ZSM-5 Phenosafranine Adsorbed onto the External Surface of the Nano-
Zeolite. J. Phys. Chem. C 2007, 111, 2310−2317. porous Host. J. Chem. Sci. 2014, 126, 945−954.
(348) Park, Y. S.; Lee, E. J.; Chun, Y. S.; Yoon, Y. D.; Yoon, K. B. (365) Easwaramoorthi, S.; Natarajan, P. Photophysical Properties of
Long-Lived Charge-Separation by Retarding Reverse Flow of Charge-
Phenosafranine (PHNS) Adsorbed on the TiO2-Incorporated Zeolite-
Balancing Cation and Zeolite-Encapsulated Ru(Bpy)32+ as Photo-
Y. Microporous Mesoporous Mater. 2005, 86, 185−190.
sensitized Electron Pump from Zeolite Framework to Externally
(366) Easwaramoorthi, S.; Natarajan, P. Characterisation and
Placed Viologen. J. Am. Chem. Soc. 2002, 124, 7123−7135.
Spectral Properties of Surface Adsorbed Phenosafranine Dye in
(349) Fukuzumi, S.; Yoshida, Y.; Urano, T.; Suenobu, T.; Imahori, H.
Zeolite-Y and ZSM-5: Photosensitisation of Embedded Nanoparticles
Extremely Slow Long-Range Electron Transfer Reactions across
of Titanium Dioxide. Microporous Mesoporous Mater. 2009, 117, 541−
Zeolite−Solution Interface. J. Am. Chem. Soc. 2001, 123, 11331−
11332. 550.
(350) Coutant, M. A.; Le, T.; Castagnola, N.; Dutta, P. K. Zeolite- (367) K, S. K.; S, C.; P, N. Photophysics and Photochemistry of
Induced Solvation Effects on Excited-State Properties of Ru(Bpy)32+: Phenosafranine Adsorbed on the Surface of ZnO Loaded Nanoporous
Implications for Intrazeolitic Photochemical Quenching Reactions. J. Materials. Dyes Pigm. 2014, 109, 206−213.
Phys. Chem. B 2000, 104, 10783−10788. (368) May, V.; Kühn, O. Charge and Energy Transfer Dynamics in
(351) Fukuzumi, S.; Urano, T.; Suenobu, T. Photoinduced Charge- Molecular Systems; Wiley-VCH: Weinheim, Germany, 2011.
Separation Using 10-Methylacridinium Ion Loaded in Zeolite Y as a (369) Hofkens, J.; Cotlet, M.; Vosch, T.; Tinnefeld, P.; Weston, K.
Photocatalyst with Negligible Back Electron Transfer across the D.; Ego, C.; Grimsdale, A.; Müllen, K.; Beljonne, D.; Brédas, J. L.; et al.
Zeolite-Solution Interface. Chem. Commun. 1996, 213−214. Revealing Competitive Förster-Type Resonance Energy-Transfer
(352) Fukuzumi, S.; Itoh, A.; Ohkubo, K.; Suenobu, T. Size-Selective Pathways in Single Bichromophoric Molecules. Proc. Natl. Acad. Sci.
Incorporation of Donor-Acceptor Linked Dyad Cations into Zeolite Y U. S. A. 2003, 100, 13146−13151.
and Long-Lived Charge Separation. RSC Adv. 2015, 5, 45582−45585. (370) Busby, M.; Kerschbaumer, H.; Calzaferri, G.; De Cola, L.
(353) Hureau, M.; Moissette, A.; Gaillard, J.; Brémard, C. Orthogonally Bifunctional Fluorescent Zeolite-L Microcrystals. Adv.
Photoinduced Electron Transfers in Zeolites: Impact of the Aluminum Mater. 2008, 20, 1614−1618.
Content on the Activation Energies. Photochem. Photobiol. Sci. 2012, (371) Inagaki, S.; Ohtani, O.; Goto, Y.; Okamoto, K.; Ikai, M.;
11, 1515−1519. Yamanaka, K.-i.; Tani, T.; Okada, T. Light Harvesting by a Periodic
(354) Spiegel, R. J.; Magrath, I. T.; Shutta, J. A. Role of Cytoplasmic Mesoporous Organosilica Chromophore. Angew. Chem., Int. Ed. 2009,
Lipids in Altering Diphenylhexatriene Fluorescence Polarization in 48, 4042−4046.
Malignant Cells. Cancer Res. 1981, 41, 452−458. (372) Beljonne, D.; Curutchet, C.; Scholes, G. D.; Silbey, R. J.
(355) Munishkina, L. A.; Fink, A. L. Fluorescence as a Method to Beyond Fö rster Resonance Energy Transfer in Biological and
Reveal Structures and Membrane-Interactions of Amyloidogenic Nanoscale Systems. J. Phys. Chem. B 2009, 113, 6583−6599.
Proteins. Biochim. Biophys. Acta, Biomembr. 2007, 1768, 1862−1885. (373) Bonacchi, S.; Genovese, D.; Juris, R.; Montalti, M.; Prodi, L.;
(356) Sánchez, J. M.; del V. Turina, A.; Perillo, M. A. Spectroscopic Rampazzo, E.; Zaccheroni, N. Luminescent Silica Nanoparticles:
Probing of Ortho-Nitrophenol Localization in Phospholipid Bilayers. J. Extending the Frontiers of Brightness. Angew. Chem., Int. Ed. 2011, 50,
Photochem. Photobiol., B 2007, 89, 56−62. 4056−4066.
(357) Moissette, A.; Hureau, M.; Col, P.; Vezin, H. Electron (374) Lõpez-Duarte, I.; Le-Quyenh, D.; Dolamic, I.; Martínez-Díaz,
Transfers in Donor-Acceptor Supramolecular Systems: Highlighting M. V.; Torres, T.; Calzaferri, G.; Brühwiler, D. On the Significance of
the Dual Donor and Acceptor Role of ZSM-5 Zeolite. J. Phys. Chem. C the Anchoring Group in the Design of Antenna Materials Based on
2016, 120, 17372−17385. Phthalocyanine Stopcocks and Zeolite L. Chem. - Eur. J. 2011, 17,
(358) Legrand, A.; Moissette, A.; Hureau, M.; Casale, S.; Massiani, P.; 1855−1862.
Vezin, H.; Mamede, A. S.; Batonneau-Gener, I. Electron Transfers in a (375) Ramachandra, S.; Popovic, Z. D.; Schuermann, K. C.;
TiO2-Containing Mor Zeolite: Synthesis of the Nanoassemblies and Cucinotta, F.; Calzaferri, G.; De Cola, L. Forster Resonance Energy

BW DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Transfer in Quantum Dot-Dye-Loaded Zeolite L Nanoassemblies. (394) Busby, M.; Blum, C.; Tibben, M.; Fibikar, S.; Calzaferri, G.;
Small 2011, 7, 1488−1494. Subramaniam, V.; De Cola, L. Time, Space, and Spectrally Resolved
(376) Calzaferri, G.; Devaux, A. Manipulation of Energy Transfer Studies on J-Aggregate Interactions in Zeolite L Nanochannels. J. Am.
Processes within the Channels of L-Zeolite. In Supramolecular Chem. Soc. 2008, 130, 10970−10976.
Photochemistry: Controlling Photochemical Processes; John Wiley & (395) Busby, M.; Devaux, A.; Blum, C.; Subramaniam, V.; Calzaferri,
Sons, Inc.: Hoboken, NJ, 2011; pp 285−387. G.; De Cola, L. Interactions of Perylene Bisimide in the One-
(377) Cucinotta, F.; Carniato, F.; Devaux, A.; De Cola, L.; Marchese, Dimensional Channels of Zeolite L. J. Phys. Chem. C 2011, 115, 5974−
L. Efficient Photoinduced Energy Transfer in a Newly Developed 5988.
Hybrid SBA-15 Photonic Antenna. Chem. - Eur. J. 2012, 18, 15310− (396) Li, P.; Wang, Y.; Li, H.; Calzaferri, G. Luminescence
15315. Enhancement after Adding Stoppers to Europium(III) Nanozeolite
(378) Mizoshita, N.; Yamanaka, K.-I.; Hiroto, S.; Shinokubo, H.; L. Angew. Chem., Int. Ed. 2014, 53, 2904−2909.
Tani, T.; Inagaki, S. Energy and Electron Transfer from Fluorescent (397) Hu, D. D.; Lin, J.; Zhang, Q.; Lu, J. N.; Wang, X. Y.; Wang, Y.
Mesostructured Organosilica Framework to Guest Dyes. Langmuir W.; Bu, F.; Ding, L. F.; Wang, L.; Wu, T. Multi-Step Host-Guest
2012, 28, 3987−3994. Energy Transfer between Inorganic Chalcogenide-Based Semiconduc-
(379) Roy, A.; Kundu, N.; Banik, D.; Sarkar, N. Comparative tor Zeolite Material and Organic Dye Molecules. Chem. Mater. 2015,
Fluorescence Resonance Energy-Transfer Study in Pluronic Triblock 27, 4099−4104.
Copolymer Micelle and Niosome Composed of Biological Compo- (398) Bruehwiler, D.; Calzaferri, G.; Torres, T.; Ramm, J. H.;
nent Cholesterol: An Investigation of Effect of Cholesterol and Gartmann, N.; Dieu, L.-Q.; Lopez-Duarte, I.; Martínez-Díaz, M. V.
Sucrose on the Fret Parameters. J. Phys. Chem. B 2016, 120, 131−142. Nanochannels for Supramolecular Organization of Luminescent
(380) Mirkovic, T.; Ostroumov, E. E.; Anna, J. M.; van Grondelle, R.; Guests. J. Mater. Chem. 2009, 19, 8040−8067.
Govindjee; Scholes, G. D. Light Absorption and Energy Transfer in (399) Lim, H.; Choi, S.-E.; Cheong, H.; Lee, J. S. Energy Transfer in
the Antenna Complexes of Photosynthetic Organisms. Chem. Rev. Dye Molecule-Containing Zeolite Monolayers. Microporous Mesopo-
2017, 117, 249−293. rous Mater. 2014, 192, 89−94.
(381) Fan, J.; Hu, M.; Zhan, P.; Peng, X. Energy Transfer Cassettes (400) Wang, L.; Liu, Y.; Chen, F.; Zhang, J.; Anpo, M. Manipulating
Based on Organic Fluorophores: Construction and Applications in Energy Transfer Processes between Rhodamine 6G and Rhodamine B
Ratiometric Sensing. Chem. Soc. Rev. 2013, 42, 29−43. in Different Mesoporous Hosts. J. Phys. Chem. C 2007, 111, 5541−
(382) Nakamura, Y.; Aratani, N.; Osuka, A. Cyclic Porphyrin Arrays 5548.
as Artificial Photosynthetic Antenna: Synthesis and Excitation Energy (401) Yamanaka, K.-i.; Okada, T.; Goto, Y.; Ikai, M.; Tani, T.;
Transfer. Chem. Soc. Rev. 2007, 36, 831−845. Inagaki, S. Dynamics of Excitation Energy Transfer from Biphenyly-
(383) Yang, J.; Yoon, M.-C.; Yoo, H.; Kim, P.; Kim, D. Excitation lene Excimers in Pore Walls of Periodic Mesoporous Organosilica to
Energy Transfer in Multiporphyrin Arrays with Cyclic Architectures: Coumarin 1 in the Mesochannels. J. Phys. Chem. C 2013, 117, 14865−
Towards Artificial Light-Harvesting Antenna Complexes. Chem. Soc. 14871.
Rev. 2012, 41, 4808−4826. (402) Trindade, F. d. J.; Triboni, E. R.; Castanheira, B.; Brochsztain,
(384) Rodríguez, H. B.; San Román, E. Effect of Concentration on S. Color-Tunable Fluorescence and White Light Emission from
the Photophysics of Dyes in Light-Scattering Materials. Photochem. Mesoporous Organosilicas Based on Energy Transfer from 1,8-
Photobiol. 2013, 89, 1273−1282. Naphthalimide Hosts to Perylenediimide Guests. J. Phys. Chem. C
(385) Cao, P.; Khorev, O.; Devaux, A.; Sägesser, L.; Kunzmann, A.; 2015, 119, 26989−26998.
Ecker, A.; Häner, R.; Brühwiler, D.; Calzaferri, G.; Belser, P. (403) Sen, T.; Jana, S.; Koner, S.; Patra, A. Energy Transfer between
Supramolecular Organization of Dye Molecules in Zeolite L Channels: Confined Dye and Surface Attached Au Nanoparticles of Mesoporous
Synthesis, Properties, and Composite Materials. Chem. - Eur. J. 2016, Silica. J. Phys. Chem. C 2010, 114, 707−714.
22, 4046−4060. (404) Sen, T.; Jana, S.; Koner, S.; Patra, A. Efficient Energy Transfer
(386) Hu, D.-D.; Wang, L.; Lin, J.; Bu, F.; Wu, T. Tuning the between Confined Dye and Y-Zeolite Functionalized Au Nano-
Efficiency of Multi-Step Energy Transfer in a Host-Guest Antenna particles. J. Phys. Chem. C 2010, 114, 19667−19672.
System Based on a Chalcogenide Semiconductor Zeolite through (405) Martin, C.; Cohen, B.; Gaamoussi, I.; Ijjaali, M.; Douhal, A.
Acidification and Solvation of Guests. J. Mater. Chem. C 2015, 3, Ultrafast Dynamics of C30 in Solution and within CDs and HSA
11747−11753. Protein. J. Phys. Chem. B 2014, 118, 5760−5771.
(387) Devaux, A.; Calzaferri, G.; Miletto, I.; Cao, P.; Belser, P.; (406) Dhir, A.; Datta, A. Shape, Size and Composition Dependence
Brühwiler, D.; Khorev, O.; Häner, R.; Kunzmann, A. Self-Absorption of Efficiency and Dynamics of Förster Resonance Energy Transfer in
and Luminescence Quantum Yields of Dye-Zeolite L Composites. J. Dye-Silica Nanoconjugates. Methods Appl. Fluoresc. 2016, 4, 024003−
Phys. Chem. C 2013, 117, 23034−23047. 024014.
(388) Calzaferri, G.; Lutkouskaya, K. Mimicking the Antenna System (407) Dhir, A.; Datta, A. FRET on Surface of Silica Nanoparticle:
of Green Plants. Photochem. Photobiol. Sci. 2008, 7, 879−910. Effect of Chromophore Concentration on Dynamics and Efficiency. J.
(389) Huber, S.; Calzaferri, G. Energy Transfer from Dye−Zeolite L Phys. Chem. C 2016, 120, 20125−20131.
Antenna Crystals to Bulk Silicon. ChemPhysChem 2004, 5, 239−242. (408) Kuroda, T.; Fujii, K.; Sakoda, K. Ultrafast Energy Transfer in a
(390) Martínez-Martínez, V.; García, R.; Gómez-Hortigüela, L.; Multichromophoric Layered Silicate. J. Phys. Chem. C 2010, 114, 983−
Pérez-Pariente, J.; López-Arbeloa, I. Modulating Dye Aggregation by 989.
Incorporation into 1D-MgAPO Nanochannels. Chem. - Eur. J. 2013, (409) Fujii, K.; Kuroda, T.; Sakoda, K.; Iyi, N. Fluorescence
19, 9859−9865. Resonance Energy Transfer and Arrangements of Fluorophores in
(391) Bujdák, J.; Martínez Martínez, V.; López Arbeloa, F.; Iyi, N. Integrated Coumarin/Cyanine Systems within Solid-State Two-
Spectral Properties of Rhodamine 3B Adsorbed on the Surface of Dimensional Nanospace. J. Photochem. Photobiol., A 2011, 225, 125−
Montmorillonites with Variable Layer Charge. Langmuir 2007, 23, 134.
1851−1859. (410) Lakowicz, J. R. Principles of Fluorescence Spectroscopy. In
(392) Bujdák, J.; Iyi, N. Molecular Aggregation of Rhodamine Dyes Principles of Fluorescence Spectroscopy; Springer: Boston, MA, 1999; pp
in Dispersions of Layered Silicates: Influence of Dye Molecular 291−319.
Structure and Silicate Properties. J. Phys. Chem. B 2006, 110, 2180− (411) Nozik, A. J.; Beard, M. C.; Luther, J. M.; Law, M.; Ellingson, R.
2186. J.; Johnson, J. C. Semiconductor Quantum Dots and Quantum Dot
(393) Calzaferri, G. Nanochannels: Hosts for the Supramolecular Arrays and Applications of Multiple Exciton Generation to Third-
Organization of Molecules and Complexes. Langmuir 2012, 28, 6216− Generation Photovoltaic Solar Cells. Chem. Rev. 2010, 110, 6873−
6231. 6890.

BX DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(412) Mackowski, S.; Wörmke, S.; Maier, A. J.; Brotosudarmo, T. H. (432) Chen, G.; Iyi, N.; Sasai, R.; Fujita, T.; Kitamura, K.
P.; Harutyunyan, H.; Hartschuh, A.; Govorov, A. O.; Scheer, H.; Intercalation of Rhodamine 6G and Oxazine 4 into Oriented Clay
Bräuchle, C. Metal-Enhanced Fluorescence of Chlorophylls in Single Films and Their Alignment. J. Mater. Res. 2002, 17, 1035−1040.
Light-Harvesting Complexes. Nano Lett. 2008, 8, 558−564. (433) Kasha, M.; Rawls, H.; Ashraf El-Bayoumi, M. The Exciton
(413) Kundu, S.; Patra, A. Nanoscale Strategies for Light Harvesting. Model in Molecular Spectroscopy. Pure Appl. Chem. 1965, 11, 371−
Chem. Rev. 2017, 117, 712−757. 392.
(414) Ma, X.; Tan, H.; Kipp, T.; Mews, A. Fluorescence (434) Eisfeld, A.; Briggs, J. S. The J- and H-Bands of Organic Dye
Enhancement, Blinking Suppression, and Gray States of Individual Aggregates. Chem. Phys. 2006, 324, 376−384.
Semiconductor Nanocrystals Close to Gold Nanoparticles. Nano Lett. (435) Davydov, A. Theory of Molecular Excitons; Springer: Boston,
2010, 10, 4166−4174. 2013.
(415) Tang, L.; Xu, J.; Guo, P.; Zhuang, X.; Tian, Y.; Wang, Y.; Duan, (436) Kobayashi, T. J-Aggregates; World Scientific: Singapore, 2012;
H.; Pan, A. Modulated Exciton-Plasmon Interactions in Au-SiO2-CdTe Vol. 2.
Composite Nanoparticles. Opt. Express 2013, 21, 11095−11100. (437) Spano, F. C. The Spectral Signatures of Frenkel Polarons in H-
(416) Zhou, N.; Liu, T.; Li, D.; Yang, D. Modifying the Fluorescence and J-Aggregates. Acc. Chem. Res. 2010, 43, 429−439.
(438) Praveen, V. K.; Ranjith, C.; Bandini, E.; Ajayaghosh, A.;
of CdTe Quantum Dots by Silica-Coated Gold Nanorods. Nanomater.
Armaroli, N. Oligo(Phenylenevinylene) Hybrids and Self-Assemblies:
Nanotechnol. 2016, 6, 16−20.
Versatile Materials for Excitation Energy Transfer. Chem. Soc. Rev.
(417) Xu, Y.; He, R.; Lin, D.; Ji, M.; Chen, J. Laser Beam Controlled
2014, 43, 4222−4242.
Drug Release from Ce6-Gold Nanorod Composites in Living Cells: A
(439) Schröter, M.; Ivanov, S. D.; Schulze, J.; Polyutov, S. P.; Yan, Y.;
Flim Study. Nanoscale 2015, 7, 2433−2441. Pullerits, T.; Kühn, O. Exciton−Vibrational Coupling in the Dynamics
(418) Brown, M. D.; Suteewong, T.; Kumar, R. S. S.; D’Innocenzo, and Spectroscopy of Frenkel Excitons in Molecular Aggregates. Phys.
V.; Petrozza, A.; Lee, M. M.; Wiesner, U.; Snaith, H. J. Plasmonic Dye- Rep. 2015, 567, 1−78.
Sensitized Solar Cells Using Core−Shell Metal−Insulator Nano- (440) Das, A.; Ghosh, S. Supramolecular Assemblies by Charge-
particles. Nano Lett. 2011, 11, 438−445. Transfer Interactions between Donor and Acceptor Chromophores.
(419) Sagarzazu, G.; Kobayashi, Y.; Murase, N.; Yang, P.; Tamai, N. Angew. Chem., Int. Ed. 2014, 53, 2038−2054.
Auger Recombination Dynamics in Hybrid Silica-Coated CdTe (441) Würthner, F.; Saha-Möller, C. R.; Fimmel, B.; Ogi, S.;
Nanocrystals. Phys. Chem. Chem. Phys. 2011, 13, 3227−3230. Leowanawat, P.; Schmidt, D. Perylene Bisimide Dye Assemblies as
(420) Sampat, S.; Karan, N. S.; Guo, T.; Htoon, H.; Hollingsworth, J. Archetype Functional Supramolecular Materials. Chem. Rev. 2016, 116,
A.; Malko, A. V. Multistate Blinking and Scaling of Recombination 962−1052.
Rates in Individual Silica-Coated CdSe/CdS Nanocrystals. ACS (442) Shim, T. K.; Lee, M. H.; Kim, D.; Kim, H. S.; Yoon, K. B.
Photonics 2015, 2, 1505−1512. Fluorescence Characteristics of Isolated Dye Molecules within
(421) Walters, R. J.; Kalkman, J.; Polman, A.; Atwater, H. A.; de Silicalite-1 Channels. J. Fluoresc. 2012, 22, 1475−1482.
Dood, M. J. A. Photoluminescence Quantum Efficiency of Dense (443) Lopez Arbeloa, F.; Tapia Estevez, M. J.; Lopez Arbeloa, T.;
Silicon Nanocrystal Ensembles in SiO2. Phys. Rev. B: Condens. Matter Lopez Arbeloa, I. Adsorption of Rhodamine 6G on Saponite. A
Mater. Phys. 2006, 73, 132302−132305. Comparative Study with Other Rhodamine 6G-Smectite Aqueous
(422) Bialas, D.; Zitzler-Kunkel, A.; Kirchner, E.; Schmidt, D.; Suspensions. Langmuir 1995, 11, 3211−3217.
Würthner, F. Structural and Quantum Chemical Analysis of Exciton (444) Vieira Ferreira, L. F.; Lemos, M. J.; Reis, M. J.; Botelho do
Coupling in Homo- and Heteroaggregate Stacks of Merocyanines. Nat. Rego, A. M. UV−Vis Absorption, Luminescence, and X-Ray
Commun. 2016, 7, 12949−12959. Photoelectron Spectroscopic Studies of Rhodamine Dyes Adsorbed
(423) Métivier, R.; Christ, T.; Kulzer, F.; Weil, T.; Müllen, K.; onto Different Pore Size Silicas. Langmuir 2000, 16, 5673−5680.
Basché, T. Single-Molecule Spectroscopy of Molecular Aggregates at (445) Chaudhuri, R.; López Arbeloa, F.; López Arbeloa, I.
Low Temperature. J. Lumin. 2004, 110, 217−224. Spectroscopic Characterization of the Adsorption of Rhodamine 3B
(424) Malfatti, L.; Kidchob, T.; Aiello, D.; Aiello, R.; Testa, F.; in Hectorite. Langmuir 2000, 16, 1285−1291.
Innocenzi, P. Aggregation States of Rhodamine 6G in Mesostructured (446) Magde, D.; Wong, R.; Seybold, P. G. Fluorescence Quantum
Silica Films. J. Phys. Chem. C 2008, 112, 16225−16230. Yields and Their Relation to Lifetimes of Rhodamine 6G and
(425) Bujdak, J.; Iyi, N.; Kaneko, Y.; Czimerova, A.; Sasai, R. Fluorescein in Nine Solvents: Improved Absolute Standards for
Molecular Arrangement of Rhodamine 6G Cations in the Films of Quantum Yields. Photochem. Photobiol. 2002, 75, 327−334.
Layered Silicates: The Effect of the Layer Charge. Phys. Chem. Chem. (447) Martínez Martínez, V. M.; López Arbeloa, F.; Vañuelos Prieto,
J.; López Arbeloa, I. Characterization of Rhodamine 6G Aggregates
Phys. 2003, 5, 4680−4685.
Intercalated in Solid Thin Films of Laponite Clay. 2 Fluorescence
(426) Martínez Martínez, V.; López Arbeloa, F.; Bañuelos Prieto, J.;
Spectroscopy. J. Phys. Chem. B 2005, 109, 7443−7450.
Arbeloa López, T.; López Arbeloa, I. Characterization of Rhodamine
(448) Carbonaro, C. M.; Orrù, F.; Ricci, P. C.; Ardu, A.; Corpino, R.;
6g Aggregates Intercalated in Solid Thin Films of Laponite Clay. 1.
Chiriu, D.; Angius, F.; Mura, A.; Cannas, C. High Efficient Fluorescent
Absorption Spectroscopy. J. Phys. Chem. B 2004, 108, 20030−20037. Stable Colloidal Sealed Dye-Doped Mesostructured Silica Nano-
(427) García, R.; Martínez-Martínez, V.; Gómez-Hortigüela, L.; particles. Microporous Mesoporous Mater. 2016, 225, 432−439.
López Arbeloa, I.;́ Pérez-Pariente, J. Anisotropic Fluorescence (449) Carbonaro, C. M.; Meinardi, F.; Ricci, P. C.; Salis, M.; Anedda,
Materials: Effect of the Synthesis Conditions over the Incorporation, A. Light Assisted Dimer to Monomer Transformation in Heavily
Alignment and Aggregation of Pyronine Y within MgAPO-5. Doped Rhodamine 6G-Porous Silica Hybrids. J. Phys. Chem. B 2009,
Microporous Mesoporous Mater. 2013, 172, 190−199. 113, 5111−5116.
(428) Calzaferri, G.; Gfeller, N. Thionine in the Cage of Zeolite L. J. (450) Synak, A.; Grobelna, B.; Kułak, L.; Lewkowicz, A.; Bojarski, P.
Phys. Chem. 1992, 96, 3428−3435. Local Dye Concentration and Spectroscopic Properties of Monomer-
(429) Xu, W.; Aydin, M.; Zakia, S.; Akins, D. L. Aggregation of Aggregate Systems in Hybrid Porous Nanolayers. J. Phys. Chem. C
Thionine within AlMCM-48. J. Phys. Chem. B 2004, 108, 5588−5593. 2015, 119, 14419−14426.
(430) Bujdák, J.; Iyi, N. Spectral and Structural Characteristics of (451) Carbonaro, C. M.; Ricci, P. C.; Grandi, S.; Marceddu, M.;
Oxazine 4/Hexadecyltrimethylammonium Montmorillonite Films. Corpino, R.; Salis, M.; Anedda, A. On the Formation of Aggregates in
Chem. Mater. 2006, 18, 2618−2624. Silica-Rhodamine 6G Type II Hybrids. RSC Adv. 2012, 2, 1905−1912.
(431) Bujdák, J.; Iyi, N. Optical Properties of Molecular Aggregates (452) Arbeloa, F. L.; Martínez; Prieto, J. B.; Arbeloa, I. L. Adsorption
of Oxazine Dyes in Dispersions of Clay Minerals. Colloid Polym. Sci. of Rhodamine 3B Dye on Saponite Colloidal Particles in Aqueous
2009, 287, 157−165. Suspensions. Langmuir 2002, 18, 2658−2664.

BY DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(453) Martin, C.; Bhattacharyya, S.; Patra, A.; Douhal, A. Single and (473) Ristanović, Z.; Hofmann, J. P.; De Cremer, G.; Kubarev, A. V.;
Multistep Energy Transfer Processes within Doped Polymer Nano- Rohnke, M.; Meirer, F.; Hofkens, J.; Roeffaers, M. B. J.; Weckhuysen,
particles. Photochem. Photobiol. Sci. 2014, 13, 1241−1252. B. M. Quantitative 3D Fluorescence Imaging of Single Catalytic
(454) Martin, C.; di Nunzio, M. R.; Cohen, B.; Douhal, A. Location Turnovers Reveals Spatiotemporal Gradients in Reactivity of Zeolite
and Freedom of Single and Double Guest in Dye-Doped Polymer H-ZSM-5 Crystals Upon Steaming. J. Am. Chem. Soc. 2015, 137,
Nanoparticles. Photochem. Photobiol. Sci. 2014, 13, 1580−1589. 6559−6568.
(455) Maiti, N. C.; Krishna, M. M. G.; Britto, P. J.; Periasamy, N. (474) Ristanović, Z.; Kubarev, A. V.; Hofkens, J.; Roeffaers, M. B. J.;
Fluorescence Dynamics of Dye Probes in Micelles. J. Phys. Chem. B Weckhuysen, B. M. Single Molecule Nanospectroscopy Visualizes
1997, 101, 11051−11060. Proton-Transfer Processes within a Zeolite Crystal. J. Am. Chem. Soc.
(456) Uppili, S.; Thomas, K. J.; Crompton, E. M.; Ramamurthy, V. 2016, 138, 13586−13596.
Probing Zeolites with Organic Molecules: Supercages of X and Y (475) Park, S. C.; Ito, T.; Higgins, D. A. Dimensionality of Diffusion
Zeolites Are Superpolar. Langmuir 2000, 16, 265−274. in Flow-Aligned Surfactant-Templated Mesoporous Silica: A Single
(457) di Nunzio, M. R.; Caballero-Mancebo, E.; Martín, C.; Cohen, Molecule Tracking Study of Pore Wall Permeability. J. Phys. Chem. C
B.; Navarro, M. T.; Corma, A.; Douhal, A. Femto-to Nanosecond 2015, 119, 26101−26110.
Photodynamics of Nile Red in Metal-Ion Exchanged Faujasites. (476) Kumarasinghe, R.; Higgins, E. D.; Ito, T.; Higgins, D. A.
Microporous Mesoporous Mater. 2018, 256, 214−226. Spectroscopic and Polarization-Dependent Single-Molecule Tracking
(458) Ferrer, M. L.; del Monte, F. Enhanced Emission of Nile Red Reveal the One-Dimensional Diffusion Pathways in Surfactant-
Fluorescent Nanoparticles Embedded in Hybrid Sol−Gel Glasses. J. Templated Mesoporous Silica. J. Phys. Chem. C 2016, 120, 715−723.
Phys. Chem. B 2005, 109, 80−86. (477) Lupo, K. M.; Hinton, D. A.; Ng, J. D.; Padilla, N. A.;
(459) Tajalli, H.; Gilani, A. G.; Zakerhamidi, M. S.; Tajalli, P. The Goldsmith, R. H. Probing Heterogeneity and Bonding at Silica
Photophysical Properties of Nile Red and Nile Blue in Ordered Surfaces through Single-Molecule Investigation of Base-Mediated
Anisotropic Media. Dyes Pigm. 2008, 78, 15−24. Linkage Failure. Langmuir 2016, 32, 9171−9179.
(460) Bardo, A. M.; Collinson, M. M.; Higgins, D. A. Nanoscale (478) Chizhik, A. M.; Tarpani, L.; Latterini, L.; Gregor, I.; Enderlein,
Properties and Matrix−Dopant Interactions in Dye-Doped Organically J.; Chizhik, A. I. Photoluminescence of a Single Quantum Emitter in a
Modified Silicate Thin Films. Chem. Mater. 2001, 13, 2713−2721. Strongly Inhomogeneous Chemical Environment. Phys. Chem. Chem.
(461) Felbeck, T.; Behnke, T.; Hoffmann, K.; Grabolle, M.; Phys. 2015, 17, 14994−15000.
Lezhnina, M. M.; Kynast, U. H.; Resch-Genger, U. Nile-Red− (479) Tarpani, L.; Ruhlandt, D.; Latterini, L.; Haehnel, D.; Gregor, I.;
Nanoclay Hybrids: Red Emissive Optical Probes for Use in Aqueous Enderlein, J.; Chizhik, A. I. Photoactivation of Luminescent Centers in
Dispersion. Langmuir 2013, 29, 11489−11497. Single SiO2 Nanoparticles. Nano Lett. 2016, 16, 4312−4316.
(462) Cser, A.; Nagy, K.; Biczók, L. Fluorescence Lifetime of Nile (480) Zicovich-Wilson, C. M.; Corma, A.; Viruela, P. Electronic
Red as a Probe for the Hydrogen Bonding Strength with Its Confinement of Molecules in Microscopic Pores - a New Cconcept
Microenvironment. Chem. Phys. Lett. 2002, 360, 473−478. Which Contributes to Explain the Catalytic Activity of Zeolites. J. Phys.
(463) Kawski, A.; Kukliński, B.; Bojarski, P. Photophysical Properties Chem. 1994, 98, 10863−10870.
(481) Zicovich-Wilson, C. M.; Planelles, J. H.; Jaskóalski, W. Spatially
and Thermochromic Shifts of Electronic Spectra of Nile Red in
Confined Simple Quantum-Mechanical Systems. Int. J. Quantum
Selected Solvents. Excited States Dipole Moments. Chem. Phys. 2009,
Chem. 1994, 50, 429−444.
359, 58−64.
(482) Zhang, L. Z.; Cheng, P. Theoretical Studies of the Electronic
(464) Marquez, F.; Garcia, H.; Palomares, E.; Fernandez, L.; Corma,
Properties of Confined Aromatic Molecules in Support of Electronic
A. Spectroscopic Evidence in Support of the Molecular Orbital
Confinement Effect. PhysChemComm 2003, 6, 62−66.
Confinement Concept: Case of Anthracene Incorporated in Zeolites. J. (483) Hyunjin, L.; Hyeonsik, C.; Sungeun, C.; Yea, N. C.; Jin, S. L.
Am. Chem. Soc. 2000, 122, 6520−6521. Optical Properties of Organic Dye Molecules Incorporated within
(465) Cordes, T.; Blum, S. A. Opportunities and Challenges in Different Zeolitic Structures. J. Korean Phys. Soc. 2011, 58, 1035−1038.
Single-Molecule and Single-Particle Fluorescence Microscopy for (484) Cohen, B.; Wang, S.; Organero, J. A.; Campo, L. F.; Sanchez,
Mechanistic Studies of Chemical Reactions. Nat. Chem. 2013, 5, F.; Douhal, A. Femtosecond Fluorescence Dynamics of a Proton-
993−999. Transfer Dye Interacting with Silica-Based Nanomaterials. J. Phys.
(466) De Cremer, G.; Sels, B. F.; De Vos, D. E.; Hofkens, J.; Chem. C 2010, 114, 6281−6289.
Roeffaers, M. B. J. Fluorescence Micro(Spectro)Scopy as a Tool to (485) Cohen, B.; Martin Á lvarez, C.; Alarcos Carmona, N.; Angel
Study Catalytic Materials in Action. Chem. Soc. Rev. 2010, 39, 4703− Organero, J.; Douhal, A. Single Molecule Photobehavior of a
4717. Chromophore Interacting with Silica-Based Nanomaterials. Phys.
(467) Higgins, D. A.; Park, S. C.; Tran-Ba, K.-H.; Ito, T. Single- Chem. Chem. Phys. 2011, 13, 1819−1826.
Molecule Investigations of Morphology and Mass Transport Dynamics (486) Weckhuysen, B. M. Porous Materials Zeolites Shine Bright.
in Nanostructured Materials. Annu. Rev. Anal. Chem. 2015, 8, 193− Nat. Mater. 2016, 15, 933−934.
216. (487) Fenwick, O.; Countiño-Gonzales, E.; Grandjean, D.; Baekelant,
(468) Li, K.; Qin, W.; Xu, Y.; Peng, T.; Li, D. Optical Approaches in W.; Richard, F.; Bonacchi, S.; De Vos, D.; Lievens, P.; Roeffaers, M.;
Study of Nanocatalysis with Single-Molecule and Single-Particle Hofkens, J.; et al. Tuning the Energetics and Tailoring the Optical
Resolution. Front. Optoelectron. 2015, 8, 379−393. Properties of Silver Clusters Confined in Zeolites. Nat. Mater. 2016,
(469) Rühle, B.; Davies, M.; Bein, T.; Bräuchle, C. Fluorescence 15, 1017−1022.
Microscopy Studies of Porous Silica Materials. Z. Naturforsch., B: J. (488) Corma, A.; García, H. Lewis Acids: From Conventional
Chem. Sci. 2013, 68, 423−444. Homogeneous to Green Homogeneous and Heterogeneous Catalysis.
(470) Sambur, J. B.; Chen, P. Approaches to Single-Nanoparticle Chem. Rev. 2003, 103, 4307−4365.
Catalysis. Annu. Rev. Phys. Chem. 2014, 65, 395−422. (489) Boronat, M.; Concepcion, P.; Corma, A.; Navarro, M. T.;
(471) Sun, X. J.; Xie, J. Y.; Xu, J. Y.; Higgins, D. A.; Hohn, K. L. Renz, M.; Valencia, S. Reactivity in the Confined Spaces of Zeolites:
Single-Molecule Studies of Acidity Distributions in Mesoporous The Interplay between Spectroscopy and Theory to Develop
Aluminosilicate Thin Films. Langmuir 2015, 31, 5667−5675. Structure-Activity Relationships for Catalysis. Phys. Chem. Chem.
(472) Ristanović, Z.; Kerssens, M. M.; Kubarev, A. V.; Hendriks, F. Phys. 2009, 11, 2876−2884.
C.; Dedecker, P.; Hofkens, J.; Roeffaers, M. B. J.; Weckhuysen, B. M. (490) Tanabe, K.; Hölderich, W. F. Industrial Application of Solid
High-Resolution Single-Molecule Fluorescence Imaging of Zeolite Acid−Base Catalysts. Appl. Catal., A 1999, 181, 399−434.
Aggregates within Real-Life Fluid Catalytic Cracking Particles. Angew. (491) Juarez, R.; Padilla, A.; Corma, A.; Garcia, H. Transition Metal
Chem., Int. Ed. 2015, 54, 1836−1840. Containing Zeolites and Mesoporous MCM-41 as Heterogeneous

BZ DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

Catalysts for the N-Alkylation of 2,4-Diaminotoluene with Dimethyl- (510) Di Fonzo, F.; Casari, C.; Russo, V.; Brunella, M.; Li Bassi, A.
carbonate. Catal. Commun. 2009, 10, 472−476. L.; Bottani, C. Hierarchically Organized Nanostructured TiO2 for
(492) Kosslick, H.; Lischke, G.; Parlitz, B.; Storek, W.; Fricke, R. Photocatalysis Applications. Nanotechnology 2009, 20, 015604−
Acidity and Active Sites of Al-MCM-41. Appl. Catal., A 1999, 184, 49− 015614.
60. (511) Anpo, M.; Kamat, P. V. Environmentally Benign Photocatalysts:
(493) Ikeue, K.; Yamashita, H.; Anpo, M.; Takewaki, T. Photo- Applications of Titanium Oxide-Based Materials; Springer-Verlag: New
catalytic Reduction of CO2 with H2O on Ti−B Zeolite Photocatalysts: York, 2010.
Effect of the Hydrophobic and Hydrophilic Properties. J. Phys. Chem. B (512) Liu, S.; Lim, M.; Amal, R. Tio2-Coated Natural Zeolite: Rapid
2001, 105, 8350−8355. Humic Acid Adsorption and Effective Photocatalytic Regeneration.
(494) Murcia-Lopez, S.; Bacariza, M. C.; Villa, K.; Lopes, J. M.; Chem. Eng. Sci. 2014, 105, 46−52.
Henriques, C.; Morante, J. R.; Andreu, T. Controlled Photocatalytic (513) Fukugaichi, S.; Henmi, T.; Matsue, N. Facile Synthesis of
Oxidation of Methane to Methanol through Surface Modification of TiO2−Zeolite Composite and Its Enhanced Photocatalytic Activity.
Beta Zeolites. ACS Catal. 2017, 7, 2878−2885. Catal. Lett. 2013, 143, 1255−1259.
(495) Ismail, A. A.; Bahnemann, D. W. Photochemical Splitting of (514) El-Roz, M.; Lakiss, L.; El Fallah, J.; Lebedev, O. I.; Thibault-
Water for Hydrogen Production by Photocatalysis: A Review. Sol. Starzyk, F.; Valtchev, V. Incorporation of Clusters of Titanium Oxide
Energy Mater. Sol. Cells 2014, 128, 85−101. in Beta Zeolite Structure by a New Cold TiCl4-Plasma Process:
(496) Dubey, N.; Rayalu, S. S.; Labhsetwar, N. K.; Devotta, S. Visible Physicochemical Properties and Photocatalytic Activity. Phys. Chem.
Light Active Zeolite-Based Photocatalysts for Hydrogen Evolution Chem. Phys. 2013, 15, 16198−16207.
from Water. Int. J. Hydrogen Energy 2008, 33, 5958−5966. (515) Salameh, C.; Nogier, J.-P.; Launay, F.; Boutros, M. Dispersion
(497) Wang, J. J.; Jing, Y. H.; Ouyang, T.; Zhang, Q.; Chang, C. T. of Colloidal TiO2 Nanoparticles on Mesoporous Materials Targeting
Photocatalytic Reduction of CO2 to Energy Products Using Cu-TiO2/ Photocatalysis Applications. Catal. Today 2015, 257, 35−40.
ZSM-5 and Co-TiO2/ZSM-5 under Low Energy Irradiation. Catal. (516) Surolia, P. K.; Jasra, R. V. Photocatalytic Degradation of p-
Commun. 2015, 59, 69−72. Nitrotoluene (PNT) Using TiO2-Modified Silver-Exchanged NaY
(498) Sastre, F.; Fornes, V.; Corma, A.; Garcia, H. Selective, Room- Zeolite: Kinetic Study and Identification of Mineralization Pathway.
Temperature Transformation of Methane to C1 Oxygenates by Deep Desalin. Water Treat. 2016, 57, 22081−22098.
UV Photolysis over Zeolites. J. Am. Chem. Soc. 2011, 133, 17257− (517) White, J. C.; Dutta, P. K. Assembly of Nanoparticles in Zeolite
17261. Y for the Photocatalytic Generation of Hydrogen from Water. J. Phys.
(499) Liu, M. M.; Hou, L. A.; Xi, B. D.; Li, Q.; Hu, X. J.; Yu, S. L. Chem. C 2011, 115, 2938−2947.
Magnetically Separable Ag/AgCl-Zero Valent Iron Particles Modified (518) Salaeh, S.; Kovacic, M.; Kosir, D.; Kusic, H.; Stangar, U. L.;
Zeolite X Heterogeneous Photocatalysts for Tetracycline Degradation Dionysiou, D. D.; Bozic, A. L. Reuse of TiO2-Based Catalyst for Solar
under Visible Light. Chem. Eng. J. 2016, 302, 475−484. Driven Water Treatment; Thermal and Chemical Reactivation. J.
(500) Gou, J. F.; Ma, Q. L.; Deng, X. Y.; Cui, Y. Q.; Zhang, H. X.; Photochem. Photobiol., A 2017, 333, 117−129.
Cheng, X. W.; Li, X. L.; Xie, M. Z.; Cheng, Q. F. Fabrication of Ag2O/ (519) Marchena, C. L.; Lerici, L.; Renzini, S.; Pierella, L.; Pizzio, L.
TiO2-Zeolite Composite and Its Enhanced Solar Light Photocatalytic Synthesis and Characterization of a Novel Tungstosilicic Acid
Performance and Mechanism for Degradation of Norfloxacin. Chem. Immobilized on Zeolites Catalyst for the Photodegradation of Methyl
Eng. J. 2017, 308, 818−826. Orange. Appl. Catal., B 2016, 188, 23−30.
(501) Jafarzadeh, A.; Sohrabnezhad, S.; Zanjanchi, M. A.; Arvand, M. (520) Wang, D. J.; Guo, R.; Wang, S. J.; Liu, F.; Wang, Y. Q.; Zhao,
Synthesis and Characterization of Thiol-Functionalized MCM-41 C. C. Synthesis and Characterization of Cobalt Phthalocyanine/MCM-
Nanofibers and Its Application as Photocatalyst. Microporous 41 and Its Photocatalytic Activity on Methyl Orange under Visible
Mesoporous Mater. 2016, 236, 109−119. Light. Desalin. Water Treat. 2016, 57, 25226−25234.
(502) Lachheb, H.; Ahmed, O.; Houas, A.; Nogier, J. P. (521) Murillo-Cremaes, N.; Lopez-Periago, A. M.; Saurina, J.; Roig,
Photocatalytic Activity of TiO2−SBA-15 under UV and Visible A.; Domingo, C. A Clean and Effective Supercritical Carbon Dioxide
Light. J. Photochem. Photobiol., A 2011, 226, 1−8. Method for the Host-Guest Synthesis and Encapsulation of Photo-
(503) Chang, F.; Wang, G.; Xie, Y.; Zhang, M.; Zhang, J.; Yang, H.-J.; active Molecules in Nanoporous Matrices. Green Chem. 2010, 12,
Hu, X. Synthesis of TiO2 Nanoparticles on Mesoporous Alumi- 2196−2204.
nosilicate Al-SBA-15 in Supercritical CO2 for Photocatalytic (522) Nezamzadeh-Ejhieh, A.; Karimi-Shamsabadi, M. Comparison
Decolorization of Methylene Blue. Ceram. Int. 2013, 39, 3823−3829. of Photocatalytic Efficiency of Supported CuO onto Micro and Nano
(504) Areerob, Y.; Cho, K. Y.; Oh, W. C. Microwave Assisted Particles of Zeolite X in Photodecolorization of Methylene Blue and
Synthesis of Graphene-Bi8La10O27-Zeolite Nanocomposite with Methyl Orange Aqueous Mixture. Appl. Catal., A 2014, 477, 83−92.
Efficient Photocatalytic Activity Towards Organic Dye Degradation. (523) Á lvaro, M.; Cabeza, J. F.; Fabuel, D.; Corma, A.; García, H.
J. Photochem. Photobiol., A 2017, 340, 157−169. Electrochemiluminescent Cells Based on Zeolite-Encapsulated Host-
(505) Ramirez-Aparicio, J.; Samaniego-Benitez, J. E.; Ramirez-Bon, R. Guest Systems: Encapsulated Ruthenium Tris-Bipyridyl. Chem. - Eur. J.
TiO2-Chabazite Semiconductor Composites for Photocatalytic Deg- 2007, 13, 3733−3738.
radation of Rhodamine under Sunlight Irradiation. Sol. Energy 2016, (524) Qu, Y.; Feng, L. J.; Liu, B. X.; Tong, C. Y.; Lu, C. L. A Facile
139, 258−265. Strategy for Synthesis of Nearly White Light Emitting Mesoporous
(506) Hu, L.; Yuan, H.; Zou, L.; Chen, F.; Hu, X. Adsorption and Silica Nanoparticles. Colloids Surf., A 2014, 441, 565−571.
Visible Light-Driven Photocatalytic Degradation of Rhodamine B in (525) Li, D.; Zhang, Y.; Fan, Z.; Chen, J.; Yu, J. Coupling of
Aqueous Solutions by Ag@AgBr/SBA-15. Appl. Surf. Sci. 2015, 355, Chromophores with Exactly Opposite Luminescence Behaviours in
706−715. Mesostructured Organosilicas for High-Efficiency Multicolour Emis-
(507) Wang, D.; Duan, Y.; Luo, Q.; Li, X.; An, J.; Bao, L.; Shi, L. sion. Chem. Sci. 2015, 6, 6097−6101.
Novel Preparation Method for a New Visible Light Photocatalyst: (526) Wang, Y.-F.; Che, J.; Zheng, Y.-C.; Zhao, Y.-Y.; Chen, F.; Jin,
Mesoporous TiO2 Supported Ag/AgBr. J. Mater. Chem. 2012, 22, S.-B.; Gong, N.-Q.; Xu, J.; Hu, Z.-B.; Liang, X.-J. Multi-Stable
4847−4854. Fluorescent Silica Nanoparticles Obtained from in Situ Doping with
(508) Linsebigler, A. L.; Lu, G.; Yates, J. T. Photocatalysis on TiO2 Aggregation-Induced Emission Molecules. J. Mater. Chem. B 2015, 3,
Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 8775−8781.
1995, 95, 735−758. (527) Kim, D.; Kim, H. S. Enhancement of Fluorescence from One-
(509) Thompson, T. L.; Yates, J. T. TiO2-Based Photocatalysis: and Two-Photon Absorption of Hemicyanine Dyes by Confinement in
Surface Defects, Oxygen and Charge Transfer. Top. Catal. 2005, 35, Silicalite-1 Nanochannels. Microporous Mesoporous Mater. 2017, 243,
197−210. 69−75.

CA DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(528) Wang, Y.; Li, H. Luminescent Materials of Zeolite Function- Lactam for Ratiometric Sensing of Lysosomal Acidity. Chem. Commun.
alized with Lanthanides. CrystEngComm 2014, 16, 9764−9778. 2011, 47, 11276−11278.
(529) Duan, T.-W.; Yan, B. Novel Luminescent Hybrids Prepared by (547) Tsou, C.-J.; Chu, C.-y.; Hung, Y.; Mou, C.-Y. A Broad Range
Incorporating a Rare Earth Ternary Complex into CdS QD Loaded Fluorescent pH Sensor Based on Hollow Mesoporous Silica
Zeolite Y Crystals through Coordination Reaction. CrystEngComm Nanoparticles, Utilising the Surface Curvature Effect. J. Mater. Chem.
2014, 16, 3395−3402. B 2013, 1, 5557−5563.
(530) Kennes, K.; Coutino-Gonzalez, E.; Martin, C.; Baekelant, W.; (548) Tsou, C.-J.; Hung, Y.; Mou, C.-Y. Hollow Mesoporous Silica
Roeffaers, M. B. J.; Van der Auweraer, M. Silver Zeolite Composites- Nanoparticles with Tunable Shell Thickness and Pore Size
Based Leds: A Novel Solid-State Lighting Approach. Adv. Funct. Mater. Distribution for Application as Broad-Ranging pH Nanosensor.
2017, 27, 1606411−1606418. Microporous Mesoporous Mater. 2014, 190, 181−188.
(531) Altantzis, T.; Coutino-Gonzalez, E.; Baekelant, W.; Martinez, (549) Wang, J. X. Mesoporous MCM-41 Embeded with Ru(II)-
G. T.; Abakumov, A. M.; Tendeloo, G. V.; Roeffaers, M. B. J.; Bals, S.; Based Chemosensor: Preparation, Characterization, and Emission
Hofkens, J. Direct Observation of Luminescent Silver Clusters Variation Towards pH. J. Lumin. 2014, 151, 41−46.
Confined in Faujasite Zeolites. ACS Nano 2016, 10, 7604−7611. (550) Vu, A.; Phillips, J.; Bühlmann, P.; Stein, A. Quenching
(532) Dong, B.; Retoux, R.; de Waele, V.; Chiodo, S. G.; Mineva, T.; Performance of Surfactant-Containing and Surfactant-Free Fluoro-
Cardin, J.; Mintova, S. Sodalite Cages of EMT Zeolite Confined phore-Doped Mesoporous Silica Films for Nitroaromatic Compound
Neutral Molecular-Like Silver Clusters. Microporous Mesoporous Mater. Detection. Chem. Mater. 2013, 25, 711−722.
2017, 244, 74−82. (551) Li, P.; Yang, D.; Li, H. Luminescence Ethylenediamine Sensor
(533) Dirin, D. N.; Protesescu, L.; Trummer, D.; Kochetygov, I. V.; Based on Terbium Complexes Entrapment. Dyes Pigm. 2016, 132,
Yakunin, S.; Krumeich, F.; Stadie, N. P.; Kovalenko, M. V. Harnessing 306−309.
Defect-Tolerance at the Nanoscale: Highly Luminescent Lead Halide (552) Li, H. R.; Cheng, W. J.; Wang, Y.; Liu, B. Y.; Zhang, W. J.;
Perovskite Nanocrystals in Mesoporous Silica Matrixes. Nano Lett. Zhang, H. J. Surface Modification and Functionalization of Micro-
2016, 16, 5866−5874. porous Hybrid Material for Luminescence Sensing. Chem. - Eur. J.
(534) Wang, H.-C.; Lin, S.-Y.; Tang, A.-C.; Singh, B. P.; Tong, H.-C.; 2010, 16, 2125−2130.
Chen, C.-Y.; Lee, Y.-C.; Tsai, T.-L.; Liu, R.-S. Mesoporous Silica (553) Sun, N. N.; Yan, B. Lanthanide Complex Inside-Outside
Particles Integrated with All-Inorganic CsPbBr3 Perovskite Quantum- Double Functionalized Zeolite a Hybrid Materials for Luminescence
Dot Nanocomposites (MP-PQDs) with High Stability and Wide Sensing. New J. Chem. 2016, 40, 6924−6930.
Color Gamut Used for Backlight Display. Angew. Chem., Int. Ed. 2016, (554) Tao, S. Y.; Li, X. B.; Wang, C.; Meng, C. G. A Dual-Emission
55, 7924−7929. Amphiphile/Dye Modified Mesoporous Silica as Fluorescent Sensor
(535) Sun, C.; Zhang, Y.; Ruan, C.; Yin, C.; Wang, X.; Wang, Y.; Yu, for the Detection of Fe3+, Cr3+ and Al3+. ChemistrySelect 2016, 1,
W. W. Efficient and Stable White LEDs with Silica-Coated Inorganic
3208−3214.
Perovskite Quantum Dots. Adv. Mater. 2016, 28, 10088−10094. (555) Coutino-Gonzalez, E.; Baekelant, W.; Grandjean, D.; Roeffaers,
(536) Huang, H.; Lin, H.; Kershaw, S. V.; Susha, A. S.; Choy, W. C.
M.; Fron, E.; Aghakhani, M. S.; Bovet, N.; Van der Auweraer, M.;
H.; Rogach, A. L. Polyhedral Oligomeric Silsesquioxane Enhances the
Lievens, P.; Vosch, T.; et al. Thermally Activated LTA(Li)-Ag Zeolites
Brightness of Perovskite Nanocrystal-Based Green Light-Emitting
with Water-Responsive Photoluminescence Properties. J. Mater. Chem.
Devices. J. Phys. Chem. Lett. 2016, 7, 4398−4404.
C 2015, 3, 11857−11867.
(537) Anaya, M.; Rubino, A.; Rojas, T. C.; Galisteo-López, J. F.;
(556) Petrizza, L.; Collot, M.; Richert, L.; Mely, Y.; Prodi, L.;
Calvo, M. E.; Míguez, H. Strong Quantum Confinement and Fast
Klymchenko, A. S. Dye-Doped Silica Nanoparticle Probes for
Photoemission Activation in CH3NH3PbI3 Perovskite Nanocrystals
Grown within Periodically Mesostructured Films. Adv. Opt. Mater. Fluorescence Lifetime Imaging of Reductive Environments in Living
2017, 5, 1601087−1601093. Cells. RSC Adv. 2016, 6, 104164−104172.
(538) Huang, S.; Li, Z.; Kong, L.; Zhu, N.; Shan, A.; Li, L. Enhancing (557) Song, C.; Ye, Z.; Wang, G.; Yuan, J.; Guan, Y. Core−Shell
the Stability of CH3NH3PbBr3 Quantum Dots by Embedding in Silica Nanoarchitectures: A Strategy to Improve the Efficiency of
Spheres Derived from Tetramethyl Orthosilicate in “Waterless” Luminescence Resonance Energy Transfer. ACS Nano 2010, 4,
Toluene. J. Am. Chem. Soc. 2016, 138, 5749−5752. 5389−5397.
(539) Malgras, V.; Henzie, J.; Takei, T.; Yamauchi, Y. Hybrid (558) Cohen, B.; Martin, C.; Iyer, S. K.; Wiesner, U.; Douhal, A.
Methylammonium Lead Halide Perovskite Nanocrystals Confined in Single Dye Molecule Behavior in Fluorescent Core-Shell Silica
Gyroidal Silica Templates. Chem. Commun. 2017, 53, 2359−2362. Nanoparticles. Chem. Mater. 2012, 24, 361−372.
(540) Vassilakopoulou, A.; Papadatos, D.; Koutselas, I. Light (559) Liu, Q.; Yang, T.; Feng, W.; Li, F. Blue-Emissive Upconversion
Emitting Diodes Based on Blends of Quasi-2D Lead Halide Nanoparticles for Low-Power-Excited Bioimaging in Vivo. J. Am.
Perovskites Stabilized within Mesoporous Silica Matrix. Microporous Chem. Soc. 2012, 134, 5390−5397.
Mesoporous Mater. 2017, 249, 165−175. (560) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H.
(541) Calzaferri, G. Artificial Photosynthesis. Top. Catal. 2010, 53, Dye-Sensitized Solar Cells. Chem. Rev. 2010, 110, 6595−6663.
130−140. (561) Á lvaro, M.; Carbonell, E.; Atienzar, P.; García, H. A Novel
(542) Rao, K. V.; Jain, A.; George, S. J. Organic-Inorganic Light- Concept for Photovoltaic Cells: Clusters of Titanium Dioxide
Harvesting Scaffolds for Luminescent Hybrids. J. Mater. Chem. C 2014, Encapsulated within Zeolites as Photoactive Semiconductors. Chem-
2, 3055−3064. PhysChem 2006, 7, 1996−2002.
(543) Jin, Y.; Kannan, S.; Wu, M.; Zhao, J. X. Toxicity of (562) Atienzar, P.; Valencia, S.; Corma, A.; García, H. Titanium-
Luminescent Silica Nanoparticles to Living Cells. Chem. Res. Toxicol. Containing Zeolites and Microporous Molecular Sieves as Photo-
2007, 20, 1126−1133. voltaic Solar Cells. ChemPhysChem 2007, 8, 1115−1119.
(544) Chen, Y.-P.; Chen, H.-A.; Hung, Y.; Chien, F.-C.; Chen, P.; (563) Atienzar, P.; Navarro, M.; Corma, A.; Garcia, H. Photovoltaic
Mou, C.-Y. Surface Charge Effect in Intracellular Localization of Activity of Ti/MCM-41. ChemPhysChem 2009, 10, 252−256.
Mesoporous Silica Nanoparticles as Probed by Fluorescent Ratio- (564) Ziółek, M.; Martín, C.; Navarro, M. T.; Garcia, H.; Douhal, A.
metric pH Imaging. RSC Adv. 2012, 2, 968−973. Confined Photodynamics of an Organic Dye for Solar Cells
(545) Doussineau, T.; Smaïhi, M.; Mohr, G. J. Two-Dye Core/Shell Encapsulated in Titanium-Doped Mesoporous Molecular Materials.
Zeolite Nanoparticles: A New Tool for Ratiometric pH Measure- J. Phys. Chem. C 2011, 115, 8858−8867.
ments. Adv. Funct. Mater. 2009, 19, 117−122. (565) Lee, Y.; Kang, M. The Optical Properties of Nanoporous
(546) Wu, S. Q.; Li, Z.; Han, J. H.; Han, S. F. Dual Colored Structured Titanium Dioxide and the Photovoltaic Efficiency on
Mesoporous Silica Nanoparticles with pH Activable Rhodamine- DSSC. Mater. Chem. Phys. 2010, 122, 284−289.

CB DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(566) Crossland, E. J. W.; Noel, N.; Sivaram, V.; Leijtens, T.; (587) Sanhai, W. R.; Sakamoto, J. H.; Canady, R.; Ferrari, M. Seven
Alexander-Webber, J. A.; Snaith, H. J. Mesoporous TiO2 Single Challenges for Nanomedicine. Nat. Nanotechnol. 2008, 3, 242−244.
Crystals Delivering Enhanced Mobility and Optoelectronic Device (588) Friberg, S.; Nyström, A. M. Nanomedicine: Will It Offer
Performance. Nature 2013, 495, 215−219. Possibilities to Overcome Multiple Drug Resistance in Cancer? J.
(567) Köppe, R.; Bossart, O.; Calzaferri, G.; Sariciftci, N. S. Advanced Nanobiotechnol. 2016, 14, 1−17.
Photon-Harvesting Concepts for Low-Energy Gap Organic Solar (589) Hare, J. I.; Lammers, T.; Ashford, M. B.; Puri, S.; Storm, G.;
Cells. Sol. Energy Mater. Sol. Cells 2007, 91, 986−995. Barry, S. T. Challenges and Strategies in Anti-Cancer Nanomedicine
(568) Kim, H. S.; Jeong, N. C.; Yoon, K. B. Photovoltaic Effects of Development: An Industry Perspective. Adv. Drug Delivery Rev. 2017,
CdS and PbS Quantum Dots Encapsulated in Zeolite Y. Langmuir 108, 25−38.
2011, 27, 14678−14688. (590) Castillo, R. R.; Colilla, M.; Vallet-Regi, M. Advances in
(569) Kim, H. S.; Yoon, K. B. Preparation and Characterization of Mesoporous Silica-Based Nanocarriers for Co-Delivery and Combi-
CdS and PbS Quantum Dots in Zeolite Y and Their Applications for nation Therapy against Cancer. Expert Opin. Drug Delivery 2017, 14,
Nonlinear Optical Materials and Solar Cell. Coord. Chem. Rev. 2014, 229−243.
263−264, 239−256. (591) Croissant, J. G.; Fatieiev, Y.; Omar, H.; Anjum, D. H.; Gurinov,
(570) Grätzel, M. The Rise of Highly Efficient and Stable Perovskite A.; Lu, J.; Tamanoi, F.; Zink, J. I.; Khashab, N. M. Periodic
Solar Cells. Acc. Chem. Res. 2017, 50, 487−491. Mesoporous Organosilica Nanoparticles with Controlled Morpholo-
(571) Malgras, V.; Tominaka, S.; Ryan, J. W.; Henzie, J.; Takei, T.; gies and High Drug/Dye Loadings for Multicargo Delivery in Cancer
Ohara, K.; Yamauchi, Y. Observation of Quantum Confinement in Cells. Chem. - Eur. J. 2016, 22, 9607−9615.
Monodisperse Methylammonium Lead Halide Perovskite Nanocryst- (592) Ruhle, B.; Saint-Cricq, P.; Zink, J. I. Externally Controlled
als Embedded in Mesoporous Silica. J. Am. Chem. Soc. 2016, 138, Nanomachines on Mesoporous Silica Nanoparticles for Biomedical
13874−13881. Applications. ChemPhysChem 2016, 17, 1769−1779.
(572) Snaith, H. J. Perovskites: The Emergence of a New Era for (593) Martinez-Carmona, M.; Colilla, M.; Vallet-Regi, M. Smart
Low-Cost, High-Efficiency Solar Cells. J. Phys. Chem. Lett. 2013, 4, Mesoporous Nanomaterials for Antitumor Therapy. Nanomaterials
3623−3630. 2015, 5, 1906−1937.
(573) Hwang, S. H.; Roh, J.; Lee, J.; Ryu, J.; Yun, J.; Jang, J. Size- (594) Sun, R.; Wang, W.; Wen, Y.; Zhang, X. Recent Advance on
Controlled SiO2 Nanoparticles as Scaffold Layers in Thin-Film Mesoporous Silica Nanoparticles-Based Controlled Release System:
Perovskite Solar Cells. J. Mater. Chem. A 2014, 2, 16429−16433. Intelligent Switches Open up New Horizon. Nanomaterials 2015, 5,
(574) Yun, J.; Ryu, J.; Lee, J.; Yu, H.; Jang, J. SiO2/TiO2 Based 2019−2053.
Hollow Nanostructures as Scaffold Layers and Al-Doping in the (595) Ansari, L.; Malaekeh-Nikouei, B. Magnetic Silica Nano-
Electron Transfer Layer for Efficient Perovskite Solar Cells. J. Mater. composites for Magnetic Hyperthermia Applications. Int. J. Hyper-
Chem. A 2016, 4, 1306−1311. thermia 2017, 33, 354−363.
(575) Lee, K.; Yoon, C.-M.; Noh, J.; Jang, J. Morphology-Controlled (596) Gisbert-Garzarán, M.; Manzano, M.; Vallet-Regí, M. pH-
Mesoporous SiO2 Nanorods for Efficient Scaffolds in Organo-Metal Responsive Mesoporous Silica and Carbon Nanoparticles for Drug
Halide Perovskite Solar Cells. Chem. Commun. 2016, 52, 4231−4234. Delivery. Bioengineering 2017, 4, 3−27.
(576) Yu, X.; Chen, S.; Yan, K.; Cai, X.; Hu, H.; Peng, M.; Chen, B.; (597) Heiden, M. G. V.; Cantley, L. C.; Thompson, C. B.
Dong, B.; Gao, X.; Zou, D. Enhanced Photovoltaic Performance of Understanding the Warburg Effect: The Metabolic Requirements of
Perovskite Solar Cells with Mesoporous SiO2 Scaffolds. J. Power Cell Proliferation. Science 2009, 324, 1029−1033.
Sources 2016, 325, 534−540. (598) Cheng, W.; Nie, J.; Xu, L.; Liang, C.; Peng, Y.; Liu, G.; Wang,
(577) Wu, R.; Yang, B.; Zhang, C.; Huang, Y.; Cui, Y.; Liu, P.; Zhou, T.; Mei, L.; Huang, L.; Zeng, X. pH-Sensitive Delivery Vehicle Based
C.; Hao, Y.; Gao, Y.; Yang, J. Prominent Efficiency Enhancement in on Folic Acid-Conjugated Polydopamine-Modified Mesoporous Silica
Perovskite Solar Cells Employing Silica-Coated Gold Nanorods. J. Nanoparticles for Targeted Cancer Therapy. ACS Appl. Mater.
Phys. Chem. C 2016, 120, 6996−7004. Interfaces 2017, 9, 18462−18473.
(578) Zhang, Y.; Zhang, Z.; Yan, W.; Zhang, B.; Feng, Y.; Asiri, A. (599) Lee, B.-Y.; Li, Z.; Clemens, D. L.; Dillon, B. J.; Hwang, A. A.;
M.; Nazeeruddin, M. K.; Gao, P. Hexagonal Mesoporous Silica Islands Zink, J. I.; Horwitz, M. A. Redox-Triggered Release of Moxifloxacin
to Enhance Photovoltaic Performance of Planar Junction Perovskite from Mesoporous Silica Nanoparticles Functionalized with Disulfide
Solar Cells. J. Mater. Chem. A 2017, 5, 1415−1420. Snap-Tops Enhances Efficacy against Pneumonic Tularemia in Mice.
(579) Nicolas, J.; Mura, S.; Brambilla, D.; Mackiewicz, N.; Couvreur, Small 2016, 12, 3690−3702.
P. Design, Functionalization Strategies and Biomedical Applications of (600) Théron, C.; Gallud, A.; Carcel, C.; Gary-Bobo, M.; Maynadier,
Targeted Biodegradable/Biocompatible Polymer-Based Nanocarriers M.; Garcia, M.; Lu, J.; Tamanoi, F.; Zink, J. I.; Man, M. W. C. Hybrid
for Drug Delivery. Chem. Soc. Rev. 2013, 42, 1147−1235. Mesoporous Silica Nanoparticles with pH -Operated and Comple-
(580) Melancon, M. P.; Zhou, M.; Li, C. Cancer Theranostics with mentary H-Bonding Caps as an Autonomous Drug-Delivery System.
near-Infrared Light-Activatable Multimodal Nanoparticles. Acc. Chem. Chem. - Eur. J. 2014, 20, 9372−9380.
Res. 2011, 44, 947−956. (601) Li, H.; Yu, H.; Zhu, C.; Hu, J.; Du, M.; Zhang, F.; Yang, D.
(581) Farokhzad, O. C.; Langer, R. Impact of Nanotechnology on Cisplatin and Doxorubicin Dual-Loaded Mesoporous Silica Nano-
Drug Delivery. ACS Nano 2009, 3, 16−20. particles for Controlled Drug Delivery. RSC Adv. 2016, 6, 94160−
(582) Wang, A. Z.; Langer, R.; Farokhzad, O. C. Nanoparticle 94169.
Delivery of Cancer Drugs. Annu. Rev. Med. 2012, 63, 185−198. (602) Datz, S.; Argyo, C.; Gattner, M.; Weiss, V.; Brunner, K.;
(583) Douhal, A. Cyclodextrin Materials Photochemistry, Photophysics Bretzler, J.; Von Schirnding, C.; Torrano, A. A.; Spada, F.; Vrabel, M.;
and Photobiology; Elsevier: Oxford, 2006; Vol. 1. et al. Genetically Designed Biomolecular Capping System for
(584) Adeoye, O.; Cabral-Marques, H. Cyclodextrin Nanosystems in Mesoporous Silica Nanoparticles Enables Receptor-Mediated Cell
Oral Drug Delivery: A Mini Review. Int. J. Pharm. 2017, 531, 521. Uptake and Controlled Drug Release. Nanoscale 2016, 8, 8101−8110.
(585) Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; (603) Karimi, M.; Sahandi Zangabad, P.; Baghaee-Ravari, S.;
Couvreur, P.; Férey, G.; Morris, R. E.; Serre, C. Metal−Organic Ghazadeh, M.; Mirshekari, H.; Hamblin, M. R. Smart Nanostructures
Frameworks in Biomedicine. Chem. Rev. 2012, 112, 1232−1268. for Cargo Delivery: Uncaging and Activating by Light. J. Am. Chem.
(586) di Nunzio, M. R.; Agostoni, V.; Cohen, B.; Gref, R.; Douhal, A. Soc. 2017, 139, 4584−4610.
A “Ship in a Bottle” Strategy to Load a Hydrophilic Anticancer Drug in (604) Linsley, C. S.; Wu, B. M. Recent Advances in Light-Responsive
Porous Metal Organic Framework Nanoparticles: Efficient Encapsu- on-Demand Drug- Delivery Systems. Ther. Delivery 2017, 8, 89−107.
lation, Matrix Stabilization, and Photodelivery. J. Med. Chem. 2014, 57, (605) Diaz-Moscoso, A.; Ballester, P. Light-Responsive Molecular
411−420. Containers. Chem. Commun. 2017, 53, 4635−4652.

CC DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews Review

(606) Yang, Y. M.; Mu, J.; Xing, B. G. Photoactivated Drug Delivery (624) Van Der Veen, R. M.; Kwon, O.-H.; Tissot, A.; Hauser, A.;
and Bioimaging. Wiley Interdiscip. Rev.: Nanomed. Nanobiotechnol. Zewail, A. H. Single-Nanoparticle Phase Transitions Visualized by
2017, 9, 1−19. Four-Dimensional Electron Microscopy. Nat. Chem. 2013, 5, 395−
(607) Noureddine, A.; Gary-Bobo, M.; Lichon, L.; Garcia, M.; Zink, 402.
J. I.; Man, M. W. C.; Cattoën, X. Bis-Clickable Mesoporous Silica (625) Yan, C.; Nishida, J.; Yuan, R.; Fayer, M. D. Water of Hydration
Nanoparticles: Straightforward Preparation of Light-Actuated Nano- Dynamics in Minerals Gypsum and Bassanite: Ultrafast 2D IR
machines for Controlled Drug Delivery with Active Targeting. Chem. - Spectroscopy of Rocks. J. Am. Chem. Soc. 2016, 138, 9694−9703.
Eur. J. 2016, 22, 9624−9630. (626) Kraack, J. P.; Hamm, P. Surface-Sensitive and Surface-Specific
(608) Liu, J. A.; Bu, W. B.; Pan, L. M.; Shi, J. L. NIR-Triggered Ultrafast Two-Dimensional Vibrational Spectroscopy. Chem. Rev.
Anticancer Drug Delivery by Upconverting Nanoparticles with 2017, 117, 10623−10664.
Integrated Azobenzene-Modified Mesoporous Silica. Angew. Chem., (627) Zanni, M. T. Two-Dimensional Infrared Spectroscopy
Int. Ed. 2013, 52, 4375−4379. Measures the Structural Dynamics of a Self-Assembled Film Only
(609) Yang, X. J.; Liu, X.; Liu, Z.; Pu, F.; Ren, J. S.; Qu, X. G. Near- One Molecule Thick. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 4890−
Infrared Light-Triggered, Targeted Drug Delivery to Cancer Cells by 4891.
Aptamer Gated Nanovehicles. Adv. Mater. 2012, 24, 2890−2895. (628) Perakis, F.; Marco, L. D.; Shalit, A.; Tang, F.; Kann, Z. R.;
(610) Yang, G. B.; Liu, J. J.; Wu, Y. F.; Feng, L. Z.; Liu, Z. Near- Kü hne, T. D.; Torre, R.; Bonn, M.; Nagata, Y. Vibrational
Infrared-Light Responsive Nanoscale Drug Delivery Systems for Spectroscopy and Dynamics of Water. Chem. Rev. 2016, 116, 7590−
Cancer Treatment. Coord. Chem. Rev. 2016, 320−321, 100−117. 7607.
(611) Croissant, J.; Chaix, A.; Mongin, O.; Wang, M.; Clement, S.; (629) Li, Z.; Mao, W.; Devadas, M. S.; Hartland, G. V. Absorption
Raehm, L.; Durand, J. O.; Hugues, V.; Blanchard-Desce, M.; Spectroscopy of Single Optically Trapped Gold Nanorods. Nano Lett.
Maynadier, M.; et al. Two-Photon-Triggered Drug Delivery Via 2015, 15, 7731−7735.
Fluorescent Nanovalves. Small 2014, 10, 1752−1755. (630) Bargheer, M.; Zhavoronkov, N.; Gritsai, Y.; Woo, J. C.; Kim, D.
(612) Croissant, J. G.; Qi, C.; Mongin, O.; Hugues, V.; Blanchard- S.; Woerner, M.; Elsaesser, T. Coherent Atomic Motions in a
Desce, M.; Raehm, L.; Cattoën, X.; Man, M. W. C.; Maynadier, M.; Nanostructure Studied by Femtosecond X-Ray Diffraction. Science
Gary-Bobo, M.; et al. Disulfide-Gated Mesoporous Silica Nano- 2004, 306, 1771−1773.
particles Designed for Two-Photon-Triggered Drug Release and (631) Chapman, H. N.; Fromme, P.; Barty, A.; White, T. A.; Kirian,
Imaging. J. Mater. Chem. B 2015, 3, 6456−6461. R. A.; Aquila, A.; Hunter, M. S.; Schulz, J.; DePonte, D. P.; Weierstall,
(613) Wang, Y.; Yang, G.; Wang, Y.; Zhao, Y.; Jiang, H.; Han, Y.; U.; et al. Femtosecond X-Ray Protein Nanocrystallography. Nature
Yang, P. Multiple Imaging and Excellent Anticancer Efficiency of an 2011, 470, 73−77.
Upconverting Nanocarrier Mediated by Single Near Infrared Light. (632) Kern, J.; Alonso-Mori, R.; Tran, R.; Hattne, J.; Gildea, R. J.;
Nanoscale 2017, 9, 4759−4769. Echols, N.; Glockner, C.; Hellmich, J.; Laksmono, H.; Sierra, R. G.;
(614) Tarn, D.; Ferris, D. P.; Barnes, J. C.; Ambrogio, M. W.; et al. Simultaneous Femtosecond X-Ray Spectroscopy and Diffraction
Stoddart, J. F.; Zink, J. I. A Reversible Light-Operated Nanovalve on of Photosystem II at Room Temperature. Science 2013, 340, 491−495.
Mesoporous Silica Nanoparticles. Nanoscale 2014, 6, 3335−3343. (633) Lindenberg, A. M.; Johnson, S. L.; Reis, D. A. Visualization of
(615) Ferris, D. P.; Zhao, Y. L.; Khashab, N. M.; Khatib, H. A.; Atomic-Scale Motions in Materials Via Femtosecond X-Ray Scattering
Stoddart, J. F.; Zink, J. I. Light-Operated Mechanized Nanoparticles. J. Techniques. Annu. Rev. Mater. Res. 2017, 47, 425−449.
Am. Chem. Soc. 2009, 131, 1686−1688. (634) Sokolowski-Tinten, K.; Blome, C.; Blums, J.; Cavalleri, A.;
(616) Dong, J. Y.; Zink, J. I. Light or Heat? The Origin of Cargo Dietrich, C.; Tarasevitch, A.; Uschmann, I.; Forster, E.; Kammler, M.;
Release from Nanoimpeller Particles Containing Upconversion Horn-Von-Hoegen, M.; et al. Femtosecond X-Ray Measurement of
Nanocrystals under IR Irradiation. Small 2015, 11, 4165−4172. Coherent Lattice Vibrations near the Lindemann Stability Limit.
(617) Deng, Y.; Huang, L.; Yang, H.; Ke, H.; He, H.; Guo, Z.; Yang, Nature 2003, 422, 287−289.
T.; Zhu, A.; Wu, H.; Chen, H. Cyanine-Anchored Silica Nanochannels (635) Yoo, B.-K.; Su, Z.; Thomas, J. M.; Zewail, A. H. On the
for Light-Driven Synergistic Thermo-Chemotherapy. Small 2017, 13, Dynamical Nature of the Active Center in a Single-Site Photocatalyst
1602747. Visualized by 4D Ultrafast Electron Microscopy. Proc. Natl. Acad. Sci.
(618) Chai, S. Q.; Guo, Y.; Zhang, Z. Y.; Chai, Z.; Ma, Y. R.; Qi, L. U. S. A. 2016, 113, 503−508.
M. Cyclodextrin-Gated Mesoporous Silica Nanoparticles as Drug
Carriers for Red Light-Induced Drug Release. Nanotechnology 2017,
28, 145101.
(619) Yang, G.; Sun, X.; Liu, J.; Feng, L.; Liu, Z. Light-Responsive,
Singlet-Oxygen-Triggered on-Demand Drug Release from Photo-
sensitizer-Doped Mesoporous Silica Nanorods for Cancer Combina-
tion Therapy. Adv. Funct. Mater. 2016, 26, 4722−4732.
(620) Vivero-Escoto, J. L.; Elnagheeb, M. Mesoporous Silica
Nanoparticles Loaded with Cisplatin and Phthalocyanine for
Combination Chemotherapy and Photodynamic Therapy in Vitro.
Nanomaterials 2015, 5, 2302−2316.
(621) Yao, X.; Chen, X.; He, C.; Chen, L.; Chen, X. Dual pH-
Responsive Mesoporous Silica Nanoparticles for Efficient Combina-
tion of Chemotherapy and Photodynamic Therapy. J. Mater. Chem. B
2015, 3, 4707−4714.
(622) Yang, Y.; Wang, Y.; Xu, W.; Zhang, X.; Shang, Y.; Xie, A.; Shen,
Y. Reduced Graphene Oxide@Mesoporous Silica-Doxorubicin/
Hydroxyapatite Inorganic Nanocomposites: Preparation and pH-
Light Dual-Triggered Synergistic Chemo-Photothermal Therapy.
Eur. J. Inorg. Chem. 2017, 2017, 2236−2246.
(623) Zhou, C.; Afonso, D.; Valetti, S.; Feiler, A.; Cardile, V.;
Graziano, A. C. E.; Conoci, S.; Sortino, S. Targeted Photodynamic
Therapy with a Folate/Sensitizer Assembly Produced from Meso-
porous Silica. Chem. - Eur. J. 2017, 23, 7672−7676.

CD DOI: 10.1021/acs.chemrev.7b00422
Chem. Rev. XXXX, XXX, XXX−XXX

You might also like