ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
Recently Viewed
You have not visited any articles yet, Please visit some articles to see contents here.
CONTENT TYPES

Figure 1Loading Img

Molecular Factors Controlling the Isomerization of Azobenzenes in the Cavity of a Flexible Coordination Cage

  • Luca Pesce
    Luca Pesce
    Department of Innovative Technologies, University of Applied Sciences and Arts of Southern Switzerland, Galleria 2, Via Cantonale 2c, CH-6928 Manno, Switzerland
    More by Luca Pesce
  • Claudio Perego
    Claudio Perego
    Department of Innovative Technologies, University of Applied Sciences and Arts of Southern Switzerland, Galleria 2, Via Cantonale 2c, CH-6928 Manno, Switzerland
  • Angela B. Grommet
    Angela B. Grommet
    Department of Organic Chemistry, Weizmann Institute of Science, Rehovot 76100, Israel
  • Rafal Klajn
    Rafal Klajn
    Department of Organic Chemistry, Weizmann Institute of Science, Rehovot 76100, Israel
    More by Rafal Klajn
  • , and 
  • Giovanni M. Pavan*
    Giovanni M. Pavan
    Department of Innovative Technologies, University of Applied Sciences and Arts of Southern Switzerland, Galleria 2, Via Cantonale 2c, CH-6928 Manno, Switzerland
    Department of Applied Science and Technology, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Turin, Italy
    *[email protected]
Cite this: J. Am. Chem. Soc. 2020, 142, 21, 9792–9802
Publication Date (Web):April 30, 2020
https://doi.org/10.1021/jacs.0c03444
Copyright © 2020 American Chemical Society
ACS AuthorChoiceACS AuthorChoicewith CC-BY-NC-NDlicense
Article Views
3945
Altmetric
-
Citations
LEARN ABOUT THESE METRICS
PDF (5 MB)
Supporting Info (1)»

Abstract

Photoswitchable molecules are employed for many applications, from the development of active materials to the design of stimuli-responsive molecular systems and light-powered molecular machines. To fully exploit their potential, we must learn ways to control the mechanism and kinetics of their photoinduced isomerization. One possible strategy involves confinement of photoresponsive switches such as azobenzenes or spiropyrans within crowded molecular environments, which may allow control over their light-induced conversion. However, the molecular factors that influence and control the switching process under realistic conditions and within dynamic molecular regimes often remain difficult to ascertain. As a case study, here we have employed molecular models to probe the isomerization of azobenzene guests within a Pd(II)-based coordination cage host in water. Atomistic molecular dynamics and metadynamics simulations allow us to characterize the flexibility of the cage in the solvent, the (rare) guest encapsulation and release events, and the relative probability/kinetics of light-induced isomerization of azobenzene analogues in these host–guest systems. In this way, we can reconstruct the mechanism of azobenzene switching inside the cage cavity and explore key molecular factors that may control this event. We obtain a molecular-level insight on the effects of crowding and host–guest interactions on azobenzene isomerization. The detailed picture elucidated by this study may enable the rational design of photoswitchable systems whose reactivity can be controlled via host–guest interactions.

Introduction

ARTICLE SECTIONS
Jump To

Photochromic molecular switches such as azobenzene,(1) spiropyran,(2,3) or arylazopyrazole(4,5) are responsive molecules that isomerize upon irradiation with light. Achieving fine control of these compounds’ isomerization kinetics is important since they represent crucial components for the development of functional photoresponsive materials,(3,6−9) light-powered molecular machines,(10−15) and in photopharmacology,(16,17) where spatiotemporal control of molecular transitions/reactions is needed. In nature, a widely employed strategy for controlling chemical reactions involves accommodating the reactants in a confined space, where encapsulation may influence and control reaction mechanism and kinetics. Enzymes, for example, can catalyze and control reactions with exquisite efficiency and fidelity by exploiting principles of molecular confinement and selective molecular/supramolecular host–guest interactions.(18−21)
In the attempt to mimic natural catalytic systems, chemists have designed synthetic cavities that exploit specific and nonspecific interactions to accommodate reactants with high selectivity in a confined space.(22) Notable examples of synthetic confined spaces are nanopores within metal–organic frameworks,(23−25) surfaces of nanoparticles(26−28) and nanopores within their aggregates,(29) microemulsion droplets,(30,31) and cavities within molecular capsules.(32) Alternatively, cavities can be found within self-assembled cages(33−38) that can encapsulate reactants in solution.(39,40) Confinement inside such cavities has been shown to be crucial in accelerating chemical reactions(29,41−43) or stabilizing otherwise unstable species.(44−48) Notable studies have reported that the photoisomerization of molecular switches can be accelerated,(49) slowed down,(50) or even completely inhibited within crowded molecular environments.(51,52) In some cases, the isomerization of photoresponsive guests may also result in expulsion of the guest(53,54) and even in disassembly of the cage.(55) A key role is attributed to the structural rigidity of synthetic host cavities, which are not capable of accommodating the large conformational changes associated with the photoisomerization of the guest.(1,3,56,57) For this reason, much interest has recently been directed toward the design of flexible molecular cages.(58−61)
In this sense, an interesting example is self-assembled metal–organic cage shown in Figure 1A.(61) This cage has been observed to encapsulate and allow for water solubilization of several different molecular switches, including azobenzene (AZB), methoxylated azobenzene (M-AZB), fluorinated azobenzene (F-AZB), various spiropyrans, as well as arylazopyrazole (AZP) (Figure 1B).(62−64) NMR and UV/vis absorption studies provided useful information about the guest encapsulation and the host−guest stoichiometry in these systems.(62−64) X-ray crystallography provided additional experimental insight on the conformations of the crystallized complexes.(62,64) Qualitatively, these efforts indicated that encapsulation and switching of the guests are enabled by large distortions within the cage structure, while confinement alters the isomerization behavior of the guests.(63)

Figure 1

Figure 1. Model host–guest systems. (A, top) Structure of the supramolecular cage studied herein, formed via the self-assembly of triimidazole-based donors and cis-blocked Pd acceptors. (A, bottom) Atomistic model of the supramolecular host cage, along with a schematic representation of its octahedral structure, which can be described by the axial and equatorial distances, D1 (red) and D2 (green). (B) Structural formulas and atomistic models of the guests studied herein: azobenzene (AZB), methoxylated azobenzene (M-AZB), fluorinated azobenzene (F-AZB), and arylazopyrazole (AZP) (here shown as trans isomers).

These findings suggest that the flexibility of the cage enables adaptation of its cavity in response to the isomerization of encapsulated guests. However, a clear understanding of the flexibility exhibited by such cages in solution is difficult to obtain experimentally. In particular, correlating such effects with isomerization events that occur on very fast time scales(52) (the transcis isomerization of excited azobenzene occurs on the order of picoseconds in the absence of confinement) and studying the mechanisms that control photoswitching under confinement are prohibitive tasks. In these cases, however, molecular simulations are extremely useful. Quantum mechanics/molecular mechanics (QM/MM), all-atom (AA), and coarse-grained (CG) simulations have been used to investigate azobenzenes and their isomerization within dense monolayers, nanocavities, supramolecular tubules, nanoparticles, and vesicles,(6,8,29,49,50,52,65−68) to name a few. Molecular models can provide detailed insight on the isomerization process(49,52) as well as on the effect of these molecular transitions on the stability(8,66,69) and out-of-equilibrium behavior of the system.(50)
Here we used AA molecular simulations to obtain an exhaustive characterization of the host–guest systems involving azo compounds AZB, M-AZB, F-AZB, and AZP (Figure 1B) as the guests and cage shown in Figure 1A as the host. Combining molecular dynamics (MD) and metadynamics (MetaD) simulations, our approach allows a thorough exploration of the flexibility of the cage in solution and of the mechanisms of guest encapsulation and release under realistic conditions. Coupled with a kinetic study of the isomerization of the excited guests within the cage, these simulations allow us to reconstruct the role of confinement on guest switching and to study the molecular factors that may allow control over this process. We found structure–reactivity relationships having a general character, which may have a profound impact on the rational design of molecular systems with programmable photoswitching properties.

Results and Discussion

ARTICLE SECTIONS
Jump To

Characterization of Cage Flexibility in Solution

The first step to understanding the behavior of the system is to determine the most favorable conformation(s) assumed by the cage, and how flexible this is in solution under realistic conditions (solvent, temperature, etc.). We started by considering the native cage, without incorporated guests. We developed an atomistic model of the cage (Figure 1A, bottom), starting from the crystal structure for the same cage containing two trans-F-AZB guests (which have been manually deleted).(62) Starting from this extended conformation of the cage, we ran a classical MD simulation in explicit water molecules at 297 K and 1 atm (see SI Methods section for details). During this MD run, we analyzed the structural rearrangements of the cage by monitoring two variables describing the geometry of the cage: D1, the axial distance between the “top” and “bottom” Pd atoms (Figure 1A, bottom; points A and B), and D2, the distance between the midpoints of the two opposite equatorial edges of the octahedral cage (points C and D). In agreement with previous studies,(63) this preliminary MD simulation showed a significant flexibility within the cage. As seen in Figure S1, D1 and D2 have been found to undergo large structural fluctuations, which, however, are observed rarely during 1 μs of a MD simulation.
To better characterize the conformational landscape and the slow structural dynamics of the cage, we turned to MetaD,(70) which allows a more exhaustive exploration of the conformational space of the cage in solution. Using the same operative conditions, we ran a MetaD simulation exploring the possible structural conformations that the cage can dynamically assume in the solvent. In the MetaD run, we used D1 and D2 as the collective variables, descriptors of the cage conformations (see SI Methods for details). From this simulation, we could reconstruct the free-energy surface (FES) for the system as a function of D1 and D2. Shown in Figure 2, the FES shows the conformational landscape of the cage (i.e., all the possible conformations that the cage can assume) associated with its relative free-energy. Dark regions in the FES identify the minimum energy (most favorable) configurations for the cage, while red, orange, yellow regions identify conformations that are increasingly higher in free-energy. We can observe a single free-energy minimum (the black region in figure) in which the cage is mildly elongated along the D1 axis (top left snapshot in Figure 2), and is generally less extended than the initial crystal structure for the cage, which is found ∼4–5 kcal/mol higher in free-energy (Figure 2, top right snapshot). Exploiting the rotation of the 12 imidazole-rings with respect to the central benzene ring,(62) the MetaD allows the system to explore a wide range of conformations (including very elongated shapes), a selection of which are shown in Figure 2. In most cases, these configurations are considerably higher in free-energy and are thus unlikely to be observed in solution. Furthermore, we calculated the probability associated with the different colored regions of the FES (see SI Methods). The result of this analysis (Figure 2B) shows that, from a statistical point of view, only the cage configurations lying within 1.5 kcal/mol of the global minimum are really accessible by the cage (dark regions on the FES and first three bins of Figure 2B). This means that the cage will most likely assume configurations within D1 ∼ 1.4–1.9 nm and D2 ∼ 0.9–1.3 nm (dark FES region). Under these conditions, these conformations will constitute ∼90% of the global population of cages in the system. It is worth adding that the shape of the FES minimum also provides important information about the flexibility of the cage in solution. In this case, the FES minimum is quite broad, which is in agreement with the experimentally observed(63) high flexibility of the cage. Altogether, this analysis provides interesting information not only on the level of flexibility of the cage under experimentally relevant conditions, but also on what we could reasonably/statistically expect to find in terms of the structural diversity of the cage in a realistic solution.

Figure 2

Figure 2. Conformational free-energy landscape of the empty cage. (A) Free-energy surface (FES) as a function of D1 (distance between the axial/red Pd atoms) and D2 (distance between the midpoints of opposite edges of the cage identified by the equatorial/green Pd atoms). The color scale in the FES indicates the free-energy associated with cage conformations on the D1D2 plane (scale and legend shown in B). Four representative snapshots are shown: the starting, extended configuration (top right) corresponding to the crystal structure of the cage,(62) the energetic minimum of the FES (top left), a D1-elongated structure (bottom right), and a D2-elongated structure (bottom left). Axial and equatorial Pd atoms are colored in red and green, respectively, while the connectivity scheme is colored in orange to facilitate interpretation of the structures. (B) Probability associated with all cage conformations as a function of the relative free-energy (bin width, 0.5 kcal/mol).

Effects of Guest Encapsulation on Cage Flexibility

The flexibility of the cage influences its ability to encapsulate guest molecules. Conversely, guest encapsulation itself may have an effect on the flexibility of the cage. For our next step, we thus studied the effect of guest encapsulation on the cage. Starting again from the crystal structure of the cage containing two trans-F-AZB guests,(62) we deleted one guest and replaced the remaining one, where necessary, in order to obtain starting models for the cage encapsulating one F-AZB, M-AZB, AZP, or AZB guest (Figure 1B). These starting host–guest complex models were then equilibrated via 1 μs of MD simulation (see SI Methods for details). In particular, we compared the effect of incorporating either trans or cis conformers of all guests, analogous to the states of the cage before and after the isomerization. In all cases, we monitored the equilibrium conformations of the cage in terms of D1 and D2. We could observe that all guests remained steadily bound within the cage during the whole MD simulation time. Figure 3 shows the equilibrium configuration for the cage in terms of D1 and D2 in all cases (colored points) on the FES of the empty cage. The colored isolines in Figure 3 identify the associated free-energy regions within 0.5 kcal/mol of the global minima for each case (the same isoline for the empty cage is shown in white).

Figure 3

Figure 3. Free-energy cost of guest encapsulation. Representative equilibrium conformations (in the D1D2 plane) of the cage encapsulating different trans (left) or cis (right) guests. For each host–guest system, we report the position of the minimum-energy conformation (colored points) and the associated isolines (same colors) enclosing all conformations within 0.5 kcal/mol from the minimum of each system. The data are projected onto the FES of the empty cage (same as Figure 2), for which we also indicate the global minimum and associated 0.5 kcal/mol isoline (in white).

On the basis of the data in Figure 3, we can draw the following conclusions. In order to incorporate a guest, the cage must undergo structural rearrangements, specifically to open up, which is accompanied by a free-energy cost. The FES data in Figure 3 allows us to assess the free-energy penalty associated with encapsulation of each guest. This value can be calculated as the difference in free-energy between the white point (empty cage) and the colored points (a higher free-energy configuration that the cage has to reach to encapsulate the guest). We have found values for ΔG in the range ∼3–5 kcal/mol, depending on the guest. Such a free-energy penalty can be seen as an entropic cost (i.e., unfavorable) for the encapsulation, as the cage elongates both axially and equatorially upon encapsulating the guest. From the data in Figure 3 we can also observe how guest encapsulation affects the flexibility of the cage. In general, the colored isolines encompass a smaller area than the white isoline associated with the empty cage. This observation suggests that the cage loses flexibility upon encapsulating a guest, the effect being more pronounced for trans (left panel) than for cis guests (right panel). Moreover, by comparing the regions occupied by the cage when hosting trans or cis guests, we note that they enclose similar free-energy values. This result suggests that isomerization of the encapsulated guest would not require a consistent free-energy cost in terms of host deformation, thus demonstrating the energetic accessibility of the process. These results are consistent with the evidence that the transcis isomerization of the guests occurs inside the cage, and the resulting cis complexes remain stable also after isomerization.(62−64)
Experiments have shown that our cage can often accommodate two trans guest molecules at the same time.(62,64) On the basis of these observations, we have modeled two additional systems wherein the cage encapsulates two trans-AZBs or two trans-F-AZBs, and have obtained equilibrated configurations for these complexes via 1 μs of MD simulation. In these cases, we observed that the cage undergoes considerable deformations compared to when only one encapsulated trans guest is present in the cavity of the model cage (see Figure S2). However, the stability observed for two-guest complexes suggests that the affinity between the encapsulated guests and the cage is significant. We anticipate that this competition between the free-energy penalty associated with the crowding inside the cage cavity and the host–guest affinity represents a crucial factor that can regulate the reactivity of these host–guest complexes.

TransCis Isomerization of Azobenzenes Inside the Cage

An interesting question is how, and to what extent, the switching of excited azobenzenes is affected by encapsulation inside the cage. The transcis isomerization of azobenzenes occurs mainly via a rotational mechanism involving the torsion of the central N–N bond, which produces the out-of-plane rotation of one end of the molecule(71,72) (see snapshots in Figure 4A). To simulate the transcis transition of the guests within the cage, we use a model for excited trans-azobenzenes (S*) that has been recently employed to study azobenzene switching inside self-assembled tubules via atomistic and coarse-grained simulations.(50) In this atomistic model, the central CNNC dihedral potential term for the azobenzene unit is modified from the black curve (unperturbed trans-azobenzene) into the blue curve in the inset of Figure 4. In this model, the trans-azobenzene guest, which is assumed to reach the excited state (S*), undergoes spontaneous transcis switching according to the correct transition pathway and kinetics during classical MD runs.(50) Starting from the equilibrated models for the cage encapsulating single molecules of each trans guest (Figure 3, left), we studied the kinetics of their switching inside and outside the cage by means of MD simulations.

Figure 4

Figure 4. Transcis transitions of azo-switches inside and outside the cage. (A) Kinetics of transcis isomerization of the excited M-AZB (top) and F-AZB (bottom) outside the cage (in solution) reported as examples. The measured transition times, reported below the isomerization arrows, are obtained from MD simulations using an atomistic model where the CNNC dihedral potential term for the trans-azobenzene derivatives (Edih, reported in the plot as a function of the dihedral angle), is changed from the black curve (native/unperturbed state) to the blue curve (excited trans-azobenzene, S*).(50) (B) Kinetics of transcis isomerization of excited M-AZB (top) and F-AZB (bottom) switches confined inside the cage. Transition times for all the guests in the cage are reported in Table 1.

As shown in Figure 4A, the isomerization of all excited unbound trans-azobenzenes studied here occurs on the time scale of picoseconds, consistent with the kinetics of isomerization for free azobenzenes (unconstrained conditions).(52) However, the process can become significantly slower as a result of encapsulation within the cage (see Figure 4B) (i.e., molecular crowding).(50) The transcis isomerization times for the guests studied herein are reported in Table 1. Interestingly, the slowing of the transition can be negligible or different by of orders of magnitude, depending on the encapsulated guest. For example, in the case of AZB and AZP, the transition kinetics is not significantly affected by the confinement in the cage. On the other hand, the switching is ∼3 or even up to ∼100 times slower for confined F-AZB and M-AZB, respectively (Figure 4B). Moreover, it is worth noting that such a considerable switching deceleration is obtained using a model where the excited trans guests cannot de-excite back.(50) Considering that the lifetime of the excited state S* of trans-azobenzene is on the order of picoseconds,(52) however, de-excitation of S* to the ground state is a non-negligible event in real systems. This suggests that the considerable deceleration in transition rates associated with encapsulated F-AZB and M-AZB could actually be, at worst, underestimated by our model. Such a difference is interesting, especially considering that cases have been reported in which azobenzene isomerization may become rare (as in highly ordered self-assembled tubules(50)) or impossible (such as within dense azobiphenyl monolayers(52)).
Table 1. Thermodynamic and Kinetic Data for trans Guest Binding and Isomerization Inside the Cage
guestΔG [kcal/mol]τoff [s]Kb [M–1]koff [s–1]kona [M–1 s–1]τtranscis [s]
AZB–5.3 ± 0.3(3.9 ± 0.2) × 10–47.87 × 1032.65 × 1032.1 × 107(1.05 ± 0.05) × 10–12
M-AZB–7.9 ± 1.3(1.2 ± 0.1) × 10–26.41 × 1058.3 × 1015.32 × 107(1.00 ± 0.05) × 10–10
F-AZB–5.3 ± 0.9(3.8 ± 0.1) × 10–37.87 × 1032.63 × 1022.1 × 106(3.0 ± 0.1) × 10–12
AZP–5.7 ± 0.8(4.2 ± 0.1) × 10–41.55 × 1042.38 × 1033.7 × 107(1.25 ± 0.05) × 10–12
a

Guest concentration in the model systems is ∼11.4 mM; to obtain the effective kon values in [s–1], the values in the table should be multiplied by 11.4 mM.

In our specific systems, guest isomerization may even result in an unstable complex. This is the case when the cage incorporates two trans-F-AZB guests at the same time. Right after isomerization is triggered in this system, one of the two encapsulated guests is expelled from the cage.(62) Interestingly, we found the same behavior in our simulations (Figure S3). Co-encapsulation of one trans- and one cis-F-AZB guest inside the cage leads to a highly unstable species, leading to the release of one of the two guests within short time scales (∼10–100 ns). The molecular factors that may affect the switching process inside the cage cavity will be discussed more in detail in the last section. Nonetheless, these results indicate that in order to fully understand and characterize the transitions in these systems, it is first necessary to study the intrinsic dynamics of the host–guest complexes and the kinetics of guest encapsulation and release.

Mechanisms of Guest Encapsulation/Release and Switching

The results discussed in the previous section provide information about the kinetics of transcis switching of excited guests inside the cage. However, this information is not sufficient to draw conclusions about whether the isomerization occurs inside or outside the cage, or about the stability of the guest encapsulation inside the host before and after the transition. To obtain a complete picture of the transition mechanism, we have studied the thermodynamics and kinetics of the guest encapsulation and expulsion in/out the cage. In fact, the ΔG for guest encapsulation (and the related kon vs koff) determines the effective probability of finding the guests inside/outside the cage and their residence time inside the cage.
Since guest encapsulation and release are, in general, rare events in these systems, it is difficult to study them via classical MD simulations. Consequently, we used MetaD simulations to investigate the thermodynamics and kinetics of guest binding/release. Starting from the equilibrated trans complexes, we conducted MetaD simulations during which the guests exchange multiple times in and out of the cage, allowing for a thorough exploration of the bound and unbound states and of the transition between them. These simulations allowed us to retrieve the free-energy difference, ΔG, between the encapsulated and free states and to calculate the corresponding Kb values (see Figure 5 and Table 1 for the data for all trans guests; the complete series including cis guests is listed in Tables S1 and S2). These data offer an exhaustive picture of the thermodynamics governing guest encapsulation, which is crucial to uncovering the probability of guest binding/release. We found that formation of the host–guest complex is energetically favored in all the tested cases, with a free-energy gain ranging from ∼3.6 kcal/mol for cis-F-AZB up to ∼8 kcal/mol for trans-M-AZB.

Figure 5

Figure 5. Thermodynamics and kinetics of trans guest binding/release. (A) Representative MetaD snapshots of the reversible binding and release of trans-M-AZB inside the cage; koff and kon denote the kinetic constants for the expulsion and encapsulation processes (the kon value inside the brackets is explicitly calculated, accounting for the guest concentration used in the model ∼11.4 mM, providing the actual rate). (B) Thermodynamic and kinetic scheme representing the expulsion and encapsulation mechanisms for trans-AZB in/out the cage as a function of the distance between the guests’ and the cage’s centers of mass (identifying IN and OUT states). (C) Thermodynamic schemes representing the expulsion/encapsulation of the trans isomers of M-AZB (green), F-AZB (cyan), and AZP (violet) guests. All ΔG differences between IN vs OUT states were computed from converged MetaD simulations, while the transition barriers were more accurately estimated from multiple infrequent MetaD runs (see the SI Methods section for further details).

As recently done to study the kinetics of monomer exchange in supramolecular polymers,(73) we then studied the kinetics of guest binding/release by means of infrequent MetaD simulations(74) (see SI Methods for details). From multiple infrequent MetaD runs activating/biasing guest release out from the cage, we could reconstruct unbiased kinetics for the event and estimate residence times for the encapsulated guest, τoff. The kinetic constant for guest release can be calculated as koff = 1/τoff. The kinetic constant for guest encapsulation (kon) can be then derived as follows: Kb = kon/koff. In this way we can obtain the complete thermodynamic and dynamic characterization of these host–guest systems, as represented in the thermodynamic schemes in Figure 5 and in the data collected in Tables 1, S1, and S2. These results reveal some subtle aspects of the binding/release processes. For example, we found that the cis conformers on average exhibit shorter residence times inside the cage with respect to the corresponding trans isomers, including more weakly binding trans isomers such as trans-AZB and trans-AZP. This effect is attributed to the lower free-energy barrier that the cis guest has to overcome to leave the cage as compared to the trans guest (see Table S2), which could be correlated with the relatively high flexibility of the host cage while accommodating cis guests (see Figure 3). Most importantly, we can observe from Table 1 that the characteristic transition times for transcis isomerization (τtrans–cis) are orders of magnitude shorter than the characteristic time for the release of trans guests from the cage (τoff). This indicates that isomerization in these systems occurs most probably inside the cage.
We also conducted MetaD simulations to study the encapsulation of a second trans guest in the case where one trans guest is already encapsulated inside the cage. We know from the experimental results(62) and from the plain MD simulations (see previous sections), that the cage can often incorporate two trans guests (e.g., F-AZB or AZB). Conversely, other molecules, such as M-AZB, can only form complexes incorporating one guest molecule.(62) The FESs obtained from MetaD simulations with two M-AZB vs two F-AZB or two AZB guests are consistent with this picture (Figure S4). In particular, these results clearly demonstrate that while incorporation of a second F-AZB, or AZB, guest in the cage is an energetically favored event (Figure S4, center and right), in the case of M-AZB, this process is highly unfavorable and unlikely (Figure S4, left). This is consistent, for example, with the available crystal structures, showing that two trans-F-AZB and two trans-AZB guests can be encapsulated within one cage, while the same cage accommodates only one trans-M-AZB at a time.(62) Moreover, while quantities such as the ΔG, Kb, kon, and koff of guest encapsulation/expulsion can be difficult to determine experimentally, the overall binding constant governing encapsulation of two trans-AZB guests within the cage could be estimated from NMR experiments, giving a value in the range of ∼109 M–2.(62) We could also estimate such overall binding constant (Kb(tot)) from the MetaD simulations using the Kb values associated with the first and the second trans-AZB guests (i.e., as Kb(tot) = Kb(guest-1) × Kb(guest-2)). The calculated value for Kb(tot) is ∼0.3 × 109 M–2 (see Table S3), which is consistent with the value obtained experimentally. The results of our simulations are also qualitatively consistent with the experimental evidence available for the other systems, for which experimental Kb values could not be obtained. We conducted infrequent MetaD simulations to compare the residence times of two F-AZB guests inside the cage before and after transcis isomerization of one guest. We found that while the switching similarly occurs within the cage, one of the two guests is then expelled promptly, within very fast time scales (nanoseconds)—orders of magnitude shorter than in the two-trans case (see Figure S3). In particular, the expulsion of cis-F-AZB is ten times more likely (or faster) than that of the trans isomer (residence times inside the cage of ∼10 ns and ∼100 ns, respectively; see also Figure S5). This finding is consistent with the experimental results(62) demonstrating that the (trans+cis)-cage ternary complex is unstable and that the isomerization of trans-F-AZB produced the rapid expulsion from the cage of one out of the two guests (most likely the cis isomer). Together, these results support the reliability of our models and provide a comprehensive characterization of the system both from thermodynamic and kinetic points of view. They also show how, in order to characterize the transition kinetics inside the cage, it is also necessary to characterize the binding and release of the guests in/out the cage.

Molecular Determinants of Guest Transitions in the Cage

In the final step, we analyzed the results from our simulations to investigate the molecular determinants that control in-cavity isomerization of the guests in these systems. Table 1 lists data describing dynamic host–guest binding in the systems and the kinetics of isomerization inside the cage cavity. Interestingly, by comparing the τtrans–cis with the τoff data, we obtain a nontrivial relationship between the residence time of the guests inside the host and the characteristic time scale needed for guest switching. In particular, we can observe that longer residence times in the cavity correspond to slower transitions (Figure 6A), suggesting that the same molecular factors regulating the stability of host–guest binding have an impact on the switching rates of the guests.

Figure 6

Figure 6. Molecular determinants of isomerization under confinement. (A) Relationship between isomerization rate (τtrans–cis) and residence times (τoff) of the guests inside the cage. (B) Relationship between τtrans–cis and potential energy of host–guest interactions, ΔEHG. (C) Relationship between τtrans–cis and the number of contacts between the cage and the guest. (D) Relationship between τtrans–cis and the volume (V) of guest molecules (see the SI Methods section for details on guest volume estimation). (E) Switching deceleration, τtrans–cis0, as a function of the increase in guest volume (%ΔVguest), in which τ0 denotes isomerization time measured at the original volume of each guest. In plots A–E, the points correspond to guests AZB (black), M-AZB (green), F-AZB (cyan), and AZP (violet). (F) Average τtrans–cis0 as a function of the average increase in guest volume (%ΔVguest), obtained by averaging all data from plot E between systems with similar %ΔVguest. The error bars indicate the standard deviation of %ΔVguest and τtrans–cis0 values. Inset: cartoon showing the volume of encapsulated M-AZB (green) inside the cage (white). The dashed lines in all plots are the logarithmic fit of the data.

In general, the energy of the host–guest interaction ΔEHG correlates with τoff. When this interaction is stronger, guest release from the cage is slower, meaning that the guest spends more time in the cage. Molecular crowding in the cavity has a similar effect. When there is a higher number of contacts between the host and the guest, τoff is longer, and guest release is less probable. We can clearly observe this effect for all simulated systems, both with trans and cis guests (see Figure S6A,B). This result is reasonable, as the interactions and the number of contacts between the host and the guest are intimately correlated to one another (see Figure S6C). This correlation is especially relevant when considering flexible hosts: stronger affinity between the host and guest leads to larger deformation of the host, as a flexible cage can structurally adapt to enhance contacts with the guest. At the same time, the relationship obtained in Figure 6A indicates that both host–guest affinity and molecular crowding impact the transition rate (τtrans–cis). This relationship between τtrans–cis and ΔEHG, and between τtrans–cis and the number of contacts, is demonstrated by the trends in Figure 6B and C, respectively. These results suggest that both host–guest affinity and molecular crowding can in principle be used to control the transition rates in the system.
Because extrapolating design principles from these plots is not trivial, we chose a more elegant strategy. We started from the simplistic consideration that, for a given impact of host–guest contacts on the switching process, the ratio between the volume occupied by the guest and the volume accessible in the cage (e.g., Vguest/Vcavity) constitutes a discriminant parameter governing transitions in the system. Given that the cage is the same in all systems, and assuming that the variations in Vcavity are negligible when comparing between the various systems (simplification), we plotted Vguest against τtrans–cis (Figure 6D) and obtained a trend similar to those described above. Even considering the approximations required to make this observation, this qualitative trend reveals a molecular-level relationship between the volume of the guest and the transition rate under confinement.
Recently, we demonstrated how reliable chemically relevant molecular models can be used as “toy models” to obtain pseudomolecular, yet useful information. By “playing” with these flexible models, we can then learn about the factors that control the system.(50,73) Having observed that guest volume has a strong impact on transition rate within the cage, we developed a computational strategy to increase the number of available data points in this trend. Starting from equilibrated models of the cage encapsulating one trans guest of each type, we artificially increased the radii of guests’ atoms in the models to increase Vguest. In all cases, the atomic radii have been increased to achieve a global increase of ∼5, 10, 20, or 30% for Vguest. Repeating the isomerization simulations for all these cases then allowed us to monitor transition deceleration as a function of the increase in guest volume (%ΔVguest). In this way, we obtained the trends shown in Figure 6E for the various guests.
These results show that transition rate generally slows down as %ΔVguest increases. Some variability between the different guests can be expected (e.g., M-AZB in green), as other features of the guest, such as chemical structure and shape, can also impact isomerization. These results are in agreement with our observation that host–guest affinity also has an impact on the rate of isomerization (Figure 6B). We also note that for some guests (M-AZB and AZP), the series is incomplete because isomerization does not occur inside the cage over a certain %ΔVguest (∼20% for AZP and ∼5–10% for M-AZB). In these cases, isomerization is either impeded by the cage or results in the release of the isomerizing guest. This observation is consistent with experimental results and with the simulations for two-guest systems. In fact, encapsulating two guests inside the cage can be assumed to be comparable to a %ΔVguest of ∼100%, well over the maximum %ΔVguest for which isomerization occurs inside the cage. As previously shown, isomerization of one guest produces rapid expulsion of one guest from two-guest systems (see Figure S3).
Finally, if we average the rate deceleration data at different values of %ΔVguest for the various systems (Figure 6E), we obtain the plot shown in Figure 6F, which shows a clear general trend between the average increase in guest volume and the deceleration of guest transitions. This trend is clearly qualitative, as other variables may also be important in controlling guest isomerization under confinement. Nevertheless, this result demonstrates a direct correlation between guest volume, crowding within the cage, and the guest transition rates. This observation helps rationalize experimental(62,75) and computational results,(50,52,76) showing that azobenzene isomerization may be hindered in highly crowded molecular systems. The generality of the computational approach presented herein offers a context for developing the idea that molecular environments can be, in principle, rationally designed to control molecular switching processes.

Conclusions

ARTICLE SECTIONS
Jump To

Learning how to control chemical reactions inside confined spaces will unlock many applications. Here we used atomistic simulations to investigate the mechanisms and molecular factors that control transcis isomerization of various azo compounds within a coordination cage. Classical molecular dynamics and metadynamics simulations allowed us to characterize these host–guest systems from thermodynamic and kinetic points of view. In this way, we obtained an exhaustive mechanistic understanding of guest encapsulation, reversible guest uptake/release dynamics and guest switching transitions, and results in agreement with experimental observations.
To elucidate the kinetics of molecular transitions under confinement, we show that it is necessary to understand the intrinsic dynamics of guest binding and release processes. We demonstrate how this equilibrium is influenced by the free-energy cost associated with host reconfiguration upon guest encapsulation, and by the energy gain due to host–guest interactions. This competition determines the residence time of encapsulated azobenzene guests inside the cage and the slowing rate of their isomerization. We identify molecular crowding and host–guest affinity as two key factors governing isomerization rate/probability in these systems. Our results demonstrate that, in principle, tuning the volume of the guest may have a direct impact on the transition rate inside the cage, or even dictate whether the transition will occur inside or outside the cage.
Overall, the advantages of the approach employed herein are many. Our approach enables a thorough characterization of (i) the flexibility and conformations accessible by the cage under given conditions of solvent and temperature, (ii) the isomerization rates for encapsulated guests, and (iii) the effect of guest encapsulation and isomerization on the flexibility and conformational distortion of the cage. The metadynamics simulations allow us to obtain molecular-level information on the thermodynamics and kinetics of the host–guest encapsulation, providing values for, e.g., kon, koff, ΔG, free-energy barriers, etc., which can be challenging to determine experimentally. From a holistic point of view, this reveals the synergies between structure, thermodynamics, and dynamics within these complex molecular systems, which concur to determine the behavior of these host−guest molecular systems. In principle, the molecular simulations approaches described herein are versatile and can be applied to investigate not only other cages but also stimuli-responsive host–guest systems in general. Our approach can provide clearer insight into the molecular factors that control structure, host–guest affinity, and dynamics, thus guiding the rational design, or customization, of systems with controllable reactivity. We envisage that this approach will find application in fields ranging from the rational design of photoresponsive host–guest systems and artificial enzymes to the control of chemical reactions in confined spaces.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.0c03444.

  • Details on the creation and parametrization of the molecular systems, simulation setup, and analysis of molecular dynamics and metadynamics simulations; additional data and figures from the simulations (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Giovanni M. Pavan - Department of Innovative Technologies, University of Applied Sciences and Arts of Southern Switzerland, Galleria 2, Via Cantonale 2c, CH-6928 Manno, SwitzerlandDepartment of Applied Science and Technology, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Turin, ItalyOrcidhttp://orcid.org/0000-0002-3473-8471 Email: [email protected]
  • Authors
    • Luca Pesce - Department of Innovative Technologies, University of Applied Sciences and Arts of Southern Switzerland, Galleria 2, Via Cantonale 2c, CH-6928 Manno, SwitzerlandOrcidhttp://orcid.org/0000-0001-6364-9577
    • Claudio Perego - Department of Innovative Technologies, University of Applied Sciences and Arts of Southern Switzerland, Galleria 2, Via Cantonale 2c, CH-6928 Manno, SwitzerlandOrcidhttp://orcid.org/0000-0001-8885-3080
    • Angela B. Grommet - Department of Organic Chemistry, Weizmann Institute of Science, Rehovot 76100, Israel
    • Rafal Klajn - Department of Organic Chemistry, Weizmann Institute of Science, Rehovot 76100, IsraelOrcidhttp://orcid.org/0000-0002-6320-8875
  • Notes

    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

GMP acknowledges the funding received by the Swiss National Science Foundation (SNSF grant number 200021_175735) and by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement no. 818776 – DYNAPOL). The authors also acknowledge the computational resources provided by the Swiss National Supercomputing Center (CSCS). RK acknowledges funding from the Minerva Foundation. ABG acknowledges funding from the Zuckerman STEM Leadership Program.

References

ARTICLE SECTIONS
Jump To

This article references 76 other publications.

  1. 1
    Bandara, H. M. D.; Burdette, S. C. Photoisomerization in Different Classes of Azobenzene. Chem. Soc. Rev. 2012, 41, 18091825,  DOI: 10.1039/C1CS15179G
  2. 2
    Shi, Z.; Peng, P.; Strohecker, D.; Liao, Y. Long-Lived Photoacid Based upon a Photochromic Reaction. J. Am. Chem. Soc. 2011, 133, 1469914703,  DOI: 10.1021/ja203851c
  3. 3
    Klajn, R. Spiropyran-Based Dynamic Materials. Chem. Soc. Rev. 2014, 43, 148184,  DOI: 10.1039/C3CS60181A
  4. 4
    Weston, C. E.; Richardson, R. D.; Haycock, P. R.; White, A. J. P.; Fuchter, M. J. Arylazopyrazoles: Azoheteroarene Photoswitches Offering Quantitative Isomerization and Long Thermal Half-Lives. J. Am. Chem. Soc. 2014, 136, 1187811881,  DOI: 10.1021/ja505444d
  5. 5
    Wang, Y.-T.; Liu, X.-Y.; Cui, G.; Fang, W.-H.; Thiel, W. Photoisomerization of Arylazopyrazole Photoswitches: Stereospecific Excited-State Relaxation. Angew. Chem., Int. Ed. 2016, 55, 1400914013,  DOI: 10.1002/anie.201607373
  6. 6
    Fredy, J. W.; Méndez-Ardoy, A.; Kwangmettatam, S.; Bochicchio, D.; Matt, B.; Stuart, M. C. A.; Huskens, J.; Katsonis, N.; Pavan, G. M.; Kudernac, T. Molecular Photoswitches Mediating the Strain-Driven Disassembly of Supramolecular Tubules. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 1185011855,  DOI: 10.1073/pnas.1711184114
  7. 7
    Yagai, S.; Iwai, K.; Yamauchi, M.; Karatsu, T.; Kitamura, A.; Uemura, S.; Morimoto, M.; Wang, H.; Würthner, F. Photocontrol Over Self-Assembled Nanostructures of π-π Stacked Dyes Supported by the Parallel Conformer of Diarylethene. Angew. Chem., Int. Ed. 2014, 53, 26022606,  DOI: 10.1002/anie.201310773
  8. 8
    Molla, M. R.; Rangadurai, P.; Antony, L.; Swaminathan, S.; de Pablo, J. J.; Thayumanavan, S. Dynamic Actuation of Glassy Polymersomes through Isomerization of a Single Azobenzene Unit at the Block Copolymer Interface. Nat. Chem. 2018, 10, 659666,  DOI: 10.1038/s41557-018-0027-6
  9. 9
    Pianowski, Z. L. Recent Implementations of Molecular Photoswitches into Smart Materials and Biological Systems. Chem. - Eur. J. 2019, 25, 51285144,  DOI: 10.1002/chem.201805814
  10. 10
    Feringa, B. L. The Art of Building Small: From Molecular Switches to Motors (Nobel Lecture). Angew. Chem., Int. Ed. 2017, 56, 1106011078,  DOI: 10.1002/anie.201702979
  11. 11
    Sauvage, J.-P. From Chemical Topology to Molecular Machines (Nobel Lecture). Angew. Chem., Int. Ed. 2017, 56, 1108011093,  DOI: 10.1002/anie.201702992
  12. 12
    Stoddart, J. F. Mechanically Interlocked Molecules (MIMs)—Molecular Shuttles, Switches, and Machines (Nobel Lecture). Angew. Chem., Int. Ed. 2017, 56, 1109411125,  DOI: 10.1002/anie.201703216
  13. 13
    Dri, C.; Peters, M. V.; Schwarz, J.; Hecht, S.; Grill, L. Spatial Periodicity in Molecular Switching. Nat. Nanotechnol. 2008, 3, 649653,  DOI: 10.1038/nnano.2008.269
  14. 14
    Kassem, S.; Lee, A. T. L.; Leigh, D. A.; Marcos, V.; Palmer, L. I.; Pisano, S. Stereodivergent Synthesis with a Programmable Molecular Machine. Nature 2017, 549, 374378,  DOI: 10.1038/nature23677
  15. 15
    Katsonis, N.; Lubomska, M.; Pollard, M. M.; Feringa, B. L.; Rudolf, P. Synthetic Light-Activated Molecular Switches and Motors on Surfaces. Prog. Surf. Sci. 2007, 82, 407434,  DOI: 10.1016/j.progsurf.2007.03.011
  16. 16
    Broichhagen, J.; Frank, J. A.; Trauner, D. A Roadmap to Success in Photopharmacology. Acc. Chem. Res. 2015, 48, 19471960,  DOI: 10.1021/acs.accounts.5b00129
  17. 17
    Weston, C. E.; Krämer, A.; Colin, F.; Yildiz, Ö.; Baud, M. G. J.; Meyer-Almes, F.-J.; Fuchter, M. J. Toward Photopharmacological Antimicrobial Chemotherapy Using Photoswitchable Amidohydrolase Inhibitors. ACS Infect. Dis. 2017, 3, 152161,  DOI: 10.1021/acsinfecdis.6b00148
  18. 18
    Fiedler, D.; Leung, D.; Bergman, R.; Raymond, K. Selective Molecular Recognition, C-H Bond Activation, and Catalysis in Nanoscale Reaction Vessels. Acc. Chem. Res. 2005, 38, 349358,  DOI: 10.1021/ar040152p
  19. 19
    Tripp, B. C.; Smith, K.; Ferry, J. G. Carbonic Anhydrase: New Insights for an Ancient Enzyme. J. Biol. Chem. 2001, 276, 4861548618,  DOI: 10.1074/jbc.R100045200
  20. 20
    Zhang, X.; Houk, K. N. Why Enzymes Are Proficient Catalysts: Beyond the Pauling Paradigm. Acc. Chem. Res. 2005, 38, 379385,  DOI: 10.1021/ar040257s
  21. 21
    Whicher, J. R.; Dutta, S.; Hansen, D. A.; Hale, W. A.; Chemler, J. A.; Dosey, A. M.; Narayan, A. R. H.; Håkansson, K.; Sherman, D. H.; Smith, J. L.; Skiniotis, G. Structural Rearrangements of a Polyketide Synthase Module during Its Catalytic Cycle. Nature 2014, 510, 560564,  DOI: 10.1038/nature13409
  22. 22
    Grommet, A. B.; Feller, M.; Klajn, R. Chemical Reactivity under Nanoconfinement. Nat. Nanotechnol. 2020, 15, 256271,  DOI: 10.1038/s41565-020-0652-2
  23. 23
    Janiak, C.; Vieth, J. K. MOFs, MILs and More: Concepts, Properties and Applications for Porous Coordination Networks (PCNs). New J. Chem. 2010, 34, 23662388,  DOI: 10.1039/c0nj00275e
  24. 24
    Goettmann, F.; Sanchez, C. How Does Confinement Affect the Catalytic Activity of Mesoporous Materials?. J. Mater. Chem. 2007, 17, 2430,  DOI: 10.1039/B608748P
  25. 25
    Boyd, P. G.; Chidambaram, A.; García-Díez, E.; Ireland, C. P.; Daff, T. D.; Bounds, R.; Gładysiak, A.; Schouwink, P.; Moosavi, S. M.; Maroto-Valer, M. M.; Reimer, J. A.; Navarro, J. A. R.; Woo, T. K.; Garcia, S.; Stylianou, K. C.; Smit, B. Data-Driven Design of Metal–Organic Frameworks for Wet Flue Gas CO2 Capture. Nature 2019, 576, 253256,  DOI: 10.1038/s41586-019-1798-7
  26. 26
    Chu, Z.; Han, Y.; Bian, T.; De, S.; Král, P.; Klajn, R. Supramolecular Control of Azobenzene Switching on Nanoparticles. J. Am. Chem. Soc. 2019, 141, 19491960,  DOI: 10.1021/jacs.8b09638
  27. 27
    Ahrens, J.; Bian, T.; Vexler, T.; Klajn, R. Irreversible Bleaching of Donor–Acceptor Stenhouse Adducts on the Surfaces of Magnetite Nanoparticles. ChemPhotoChem. 2017, 1, 230236,  DOI: 10.1002/cptc.201700009
  28. 28
    Zdobinsky, T.; Sankar Maiti, P.; Klajn, R. Support Curvature and Conformational Freedom Control Chemical Reactivity of Immobilized Species. J. Am. Chem. Soc. 2014, 136, 27112714,  DOI: 10.1021/ja411573a
  29. 29
    Zhao, H.; Sen, S.; Udayabhaskararao, T.; Sawczyk, M.; Kučanda, K.; Manna, D.; Kundu, P. K.; Lee, J.-W.; Král, P.; Klajn, R. Reversible Trapping and Reaction Acceleration within Dynamically Self-Assembling Nanoflasks. Nat. Nanotechnol. 2016, 11, 8288,  DOI: 10.1038/nnano.2015.256
  30. 30
    Fallah-Araghi, A.; Meguellati, K.; Baret, J.-C.; Harrak, A. E.; Mangeat, T.; Karplus, M.; Ladame, S.; Marques, C. M.; Griffiths, A. D. Enhanced Chemical Synthesis at Soft Interfaces: A Universal Reaction-Adsorption Mechanism in Microcompartments. Phys. Rev. Lett. 2014, 112, 028301  DOI: 10.1103/PhysRevLett.112.028301
  31. 31
    Franco, C.; Rodríguez-San-Miguel, D.; Sorrenti, A.; Sevim, S.; Pons, R.; Platero-Prats, A. E.; Pavlovic, M.; Szilágyi, I.; Ruiz Gonzalez, M. L.; González-Calbet, J. M.; Bochicchio, D.; Pesce, L.; Pavan, G. M.; Imaz, I.; Cano-Sarabia, M.; Maspoch, D.; Pané, S.; de Mello, A. J.; Zamora, F.; Puigmartí-Luis, J. Biomimetic Synthesis of Sub-20 nm Covalent Organic Frameworks in Water. J. Am. Chem. Soc. 2020, 142, 35403547,  DOI: 10.1021/jacs.9b12389
  32. 32
    Zhang, G.; Mastalerz, M. Organic Cage Compounds – from Shape-Persistency to Function. Chem. Soc. Rev. 2014, 43, 19341947,  DOI: 10.1039/C3CS60358J
  33. 33
    Kang, J.; Rebek, J. Acceleration of a Diels–Alder Reaction by a Self-Assembled Molecular Capsule. Nature 1997, 385, 5052,  DOI: 10.1038/385050a0
  34. 34
    Yoshizawa, M.; Klosterman, J. K.; Fujita, M. Functional Molecular Flasks: New Properties and Reactions within Discrete, Self-Assembled Hosts. Angew. Chem., Int. Ed. 2009, 48, 34183438,  DOI: 10.1002/anie.200805340
  35. 35
    Roy, B.; Ghosh, A. K.; Srivastava, S.; D’Silva, P.; Mukherjee, P. S. A Pd8 Tetrafacial Molecular Barrel as Carrier for Water Insoluble Fluorophore. J. Am. Chem. Soc. 2015, 137, 1191611919,  DOI: 10.1021/jacs.5b08008
  36. 36
    Wang, K.; Cai, X.; Yao, W.; Tang, D.; Kataria, R.; Ashbaugh, H. S.; Byers, L. D.; Gibb, B. C. Electrostatic Control of Macrocyclization Reactions within Nanospaces. J. Am. Chem. Soc. 2019, 141, 67406747,  DOI: 10.1021/jacs.9b02287
  37. 37
    Cook, T. R.; Stang, P. J. Recent Developments in the Preparation and Chemistry of Metallacycles and Metallacages via Coordination. Chem. Rev. 2015, 115, 70017045,  DOI: 10.1021/cr5005666
  38. 38
    Sepehrpour, H.; Fu, W.; Sun, Y.; Stang, P. J. Biomedically Relevant Self-Assembled Metallacycles and Metallacages. J. Am. Chem. Soc. 2019, 141, 1400514020,  DOI: 10.1021/jacs.9b06222
  39. 39
    Liu, M.; Zhang, L.; Little, M. A.; Kapil, V.; Ceriotti, M.; Yang, S.; Ding, L.; Holden, D. L.; Balderas-Xicohténcatl, R.; He, D.; Clowes, R.; Chong, S. Y.; Schütz, G.; Chen, L.; Hirscher, M.; Cooper, A. I. Barely Porous Organic Cages for Hydrogen Isotope Separation. Science 2019, 366, 613620,  DOI: 10.1126/science.aax7427
  40. 40
    Merget, S.; Catti, L.; Piccini, G.; Tiefenbacher, K. Requirements for Terpene Cyclizations inside the Supramolecular Resorcinarene Capsule: Bound Water and Its Protonation Determine the Catalytic Activity. J. Am. Chem. Soc. 2020, 142, 44004410,  DOI: 10.1021/jacs.9b13239
  41. 41
    Yoshizawa, M.; Tamura, M.; Fujita, M. Diels-Alder in Aqueous Molecular Hosts: Unusual Regioselectivity and Efficient Catalysis. Science 2006, 312, 251254,  DOI: 10.1126/science.1124985
  42. 42
    Ueda, Y.; Ito, H.; Fujita, D.; Fujita, M. Permeable Self-Assembled Molecular Containers for Catalyst Isolation Enabling Two-Step Cascade Reactions. J. Am. Chem. Soc. 2017, 139, 60906093,  DOI: 10.1021/jacs.7b02745
  43. 43
    Maestri, M.; Iglesia, E. First-Principles Theoretical Assessment of Catalysis by Confinement: NO–O2 Reactions within Voids of Molecular Dimensions in Siliceous Crystalline Frameworks. Phys. Chem. Chem. Phys. 2018, 20, 1572515735,  DOI: 10.1039/C8CP01615A
  44. 44
    Mal, P.; Breiner, B.; Rissanen, K.; Nitschke, J. R. White Phosphorus Is Air-Stable Within a Self-Assembled Tetrahedral Capsule. Science 2009, 324, 1697,  DOI: 10.1126/science.1175313
  45. 45
    Yamashina, M.; Sei, Y.; Akita, M.; Yoshizawa, M. Safe Storage of Radical Initiators within a Polyaromatic Nanocapsule. Nat. Commun. 2014, 5, 4662,  DOI: 10.1038/ncomms5662
  46. 46
    Galan, A.; Ballester, P. Stabilization of Reactive Species by Supramolecular Encapsulation. Chem. Soc. Rev. 2016, 45, 17201737,  DOI: 10.1039/C5CS00861A
  47. 47
    Qiu, Y.; Antony, L. W.; Torkelson, J. M.; de Pablo, J. J.; Ediger, M. D. Tenfold Increase in the Photostability of an Azobenzene Guest in Vapor-Deposited Glass Mixtures. J. Chem. Phys. 2018, 149, 204503,  DOI: 10.1063/1.5052003
  48. 48
    Qiu, Y.; Antony, L. W.; de Pablo, J. J.; Ediger, M. D. Photostability Can Be Significantly Modulated by Molecular Packing in Glasses. J. Am. Chem. Soc. 2016, 138, 1128211289,  DOI: 10.1021/jacs.6b06372
  49. 49
    Fregoni, J.; Granucci, G.; Persico, M.; Corni, S. Strong Coupling with Light Enhances the Photoisomerization Quantum Yield of Azobenzene. Chem. 2020, 6, 250265,  DOI: 10.1016/j.chempr.2019.11.001
  50. 50
    Bochicchio, D.; Kwangmettatam, S.; Kudernac, T.; Pavan, G. M. How Defects Control the Out-of-Equilibrium Dissipative Evolution of a Supramolecular Tubule. ACS Nano 2019, 13, 43224334,  DOI: 10.1021/acsnano.8b09523
  51. 51
    Kusukawa, T.; Fujita, M. Ship-in-a-Bottle” Formation of Stable Hydrophobic Dimers of Cis-Azobenzene and -Stilbene Derivatives in a Self-Assembled Coordination Nanocage. J. Am. Chem. Soc. 1999, 121, 13971398,  DOI: 10.1021/ja9837295
  52. 52
    Cantatore, V.; Granucci, G.; Rousseau, G.; Padula, G.; Persico, M. Photoisomerization of Self-Assembled Monolayers of Azobiphenyls: Simulations Highlight the Role of Packing and Defects. J. Phys. Chem. Lett. 2016, 7, 40274031,  DOI: 10.1021/acs.jpclett.6b02018
  53. 53
    Clever, G. H.; Tashiro, S.; Shionoya, M. Light-Triggered Crystallization of a Molecular Host-Guest Complex. J. Am. Chem. Soc. 2010, 132, 99739975,  DOI: 10.1021/ja103620z
  54. 54
    Dube, H.; Ajami, D.; Rebek, J. Photochemical Control of Reversible Encapsulation. Angew. Chem., Int. Ed. 2010, 49, 31923195,  DOI: 10.1002/anie.201000876
  55. 55
    Mohan Raj, A.; Raymo, F. M.; Ramamurthy, V. Reversible Disassembly–Assembly of Octa Acid–Guest Capsule in Water Triggered by a Photochromic Process. Org. Lett. 2016, 18, 15661569,  DOI: 10.1021/acs.orglett.6b00405
  56. 56
    Yang, Y.; Hughes, R. P.; Aprahamian, I. Visible Light Switching of a BF2-Coordinated Azo Compound. J. Am. Chem. Soc. 2012, 134, 1522115224,  DOI: 10.1021/ja306030d
  57. 57
    Helmy, S.; Leibfarth, F. A.; Oh, S.; Poelma, J. E.; Hawker, C. J.; Read de Alaniz, J. Photoswitching Using Visible Light: A New Class of Organic Photochromic Molecules. J. Am. Chem. Soc. 2014, 136, 81698172,  DOI: 10.1021/ja503016b
  58. 58
    Zhang, D.; Ronson, T. K.; Mosquera, J.; Martinez, A.; Guy, L.; Nitschke, J. R. Anion Binding in Water Drives Structural Adaptation in an Azaphosphatrane-Functionalized FeII4L4 Tetrahedron. J. Am. Chem. Soc. 2017, 139, 65746577,  DOI: 10.1021/jacs.7b02950
  59. 59
    Rizzuto, F. J.; Nitschke, J. R. Stereochemical Plasticity Modulates Cooperative Binding in a CoII12L6 Cuboctahedron. Nat. Chem. 2017, 9, 903908,  DOI: 10.1038/nchem.2758
  60. 60
    Mondal, P.; Sarkar, S.; Rath, S. P. Cyclic Bis-Porphyrin-Based Flexible Molecular Containers: Controlling Guest Arrangements and Supramolecular Catalysis by Tuning Cavity Size. Chem. - Eur. J. 2017, 23, 70937103,  DOI: 10.1002/chem.201700577
  61. 61
    Samanta, D.; Mukherjee, S.; Patil, Y. P.; Mukherjee, P. S. Self-Assembled Pd6 Open Cage with Triimidazole Walls and the Use of Its Confined Nanospace for Catalytic Knoevenagel- and Diels–Alder Reactions in Aqueous Medium. Chem. - Eur. J. 2012, 18, 1232212329,  DOI: 10.1002/chem.201201679
  62. 62
    Samanta, D.; Gemen, J.; Chu, Z.; Diskin-Posner, Y.; Shimon, L. J. W.; Klajn, R. Reversible Photoswitching of Encapsulated Azobenzenes in Water. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 93799384,  DOI: 10.1073/pnas.1712787115
  63. 63
    Samanta, D.; Galaktionova, D.; Gemen, J.; Shimon, L. J. W.; Diskin-Posner, Y.; Avram, L.; Král, P.; Klajn, R. Reversible Chromism of Spiropyran in the Cavity of a Flexible Coordination Cage. Nat. Commun. 2018, 9, 641,  DOI: 10.1038/s41467-017-02715-6
  64. 64
    Hanopolskyi, A. I.; De, S.; Białek, M. J.; Diskin-Posner, Y.; Avram, L.; Feller, M.; Klajn, R. Reversible Switching of Arylazopyrazole within a Metal–Organic Cage. Beilstein J. Org. Chem. 2019, 15, 23982407,  DOI: 10.3762/bjoc.15.232
  65. 65
    Böckmann, M.; Peter, C.; Site, L. D.; Doltsinis, N. L.; Kremer, K.; Marx, D. Atomistic Force Field for Azobenzene Compounds Adapted for QM/MM Simulations with Applications to Liquids and Liquid Crystals. J. Chem. Theory Comput. 2007, 3, 17891802,  DOI: 10.1021/ct7000733
  66. 66
    Peter, C.; Site, L. D.; Kremer, K. Classical Simulations from the Atomistic to the Mesoscale and Back: Coarse Graining an Azobenzene Liquid Crystal. Soft Matter 2008, 4, 859869,  DOI: 10.1039/b717324e
  67. 67
    Ilnytskyi, J. M.; Slyusarchuk, A.; Saphiannikova, M. Photocontrollable Self-Assembly of Azobenzene-Decorated Nanoparticles in Bulk: Computer Simulation Study. Macromolecules 2016, 49, 92729282,  DOI: 10.1021/acs.macromol.6b01871
  68. 68
    Osella, S.; Minoia, A.; Beljonne, D. Combined Molecular Dynamics and Density Functional Theory Study of Azobenzene–Graphene Interfaces. J. Phys. Chem. C 2016, 120, 66516658,  DOI: 10.1021/acs.jpcc.6b00393
  69. 69
    Döbbelin, M.; Ciesielski, A.; Haar, S.; Osella, S.; Bruna, M.; Minoia, A.; Grisanti, L.; Mosciatti, T.; Richard, F.; Prasetyanto, E. A.; De Cola, L.; Palermo, V.; Mazzaro, R.; Morandi, V.; Lazzaroni, R.; Ferrari, A. C.; Beljonne, D.; Samorì, P. Light-Enhanced Liquid-Phase Exfoliation and Current Photoswitching in Graphene–Azobenzene Composites. Nat. Commun. 2016, 7, 11090,  DOI: 10.1038/ncomms11090
  70. 70
    Laio, A.; Parrinello, M. Escaping Free-Energy Minima. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 1256212566,  DOI: 10.1073/pnas.202427399
  71. 71
    Pederzoli, M.; Pittner, J.; Barbatti, M.; Lischka, H. Nonadiabatic Molecular Dynamics Study of the CisTrans Photoisomerization of Azobenzene Excited to the S1 State. J. Phys. Chem. A 2011, 115, 1113611143,  DOI: 10.1021/jp2013094
  72. 72
    Tiago, M. L.; Ismail-Beigi, S.; Louie, S. G. Photoisomerization of Azobenzene from First-Principles Constrained Density-Functional Calculations. J. Chem. Phys. 2005, 122, 094311  DOI: 10.1063/1.1861873
  73. 73
    Bochicchio, D.; Salvalaglio, M.; Pavan, G. M. Into the Dynamics of a Supramolecular Polymer at Submolecular Resolution. Nat. Commun. 2017, 8, 147,  DOI: 10.1038/s41467-017-00189-0
  74. 74
    Tiwary, P.; Parrinello, M. From Metadynamics to Dynamics. Phys. Rev. Lett. 2013, 111, 230602,  DOI: 10.1103/PhysRevLett.111.230602
  75. 75
    Pace, G.; Ferri, V.; Grave, C.; Elbing, M.; von Hänisch, C.; Zharnikov, M.; Mayor, M.; Rampi, M. A.; Samorì, P. Cooperative Light-Induced Molecular Movements of Highly Ordered Azobenzene Self-Assembled Monolayers. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 9937,  DOI: 10.1073/pnas.0703748104
  76. 76
    Titov, E.; Granucci, G.; Götze, J. P.; Persico, M.; Saalfrank, P. Dynamics of Azobenzene Dimer Photoisomerization: Electronic and Steric Effects. J. Phys. Chem. Lett. 2016, 7, 35913596,  DOI: 10.1021/acs.jpclett.6b01401

Cited By


This article is cited by 21 publications.

  1. Sayan Sarkar, Piyali Sarkar, Pradyut Ghosh. Heteroditopic Macrobicyclic Molecular Vessels for Single Step Aerial Oxidative Transformation of Primary Alcohol Appended Cross Azobenzenes. The Journal of Organic Chemistry 2021, 86 (9) , 6648-6664. https://doi.org/10.1021/acs.joc.1c00409
  2. Zhao-Tao Shi, Yi-Xiong Hu, Zhubin Hu, Qi Zhang, Shao-Yu Chen, Meng Chen, Jing-Jing Yu, Guang-Qiang Yin, Haitao Sun, Lin Xu, Xiaopeng Li, Ben L. Feringa, Hai-Bo Yang, He Tian, Da-Hui Qu. Visible-Light-Driven Rotation of Molecular Motors in Discrete Supramolecular Metallacycles. Journal of the American Chemical Society 2021, 143 (1) , 442-452. https://doi.org/10.1021/jacs.0c11752
  3. Charly Empereur-Mot, Luca Pesce, Giovanni Doni, Davide Bochicchio, Riccardo Capelli, Claudio Perego, Giovanni M. Pavan. Swarm-CG: Automatic Parametrization of Bonded Terms in MARTINI-Based Coarse-Grained Models of Simple to Complex Molecules via Fuzzy Self-Tuning Particle Swarm Optimization. ACS Omega 2020, 5 (50) , 32823-32843. https://doi.org/10.1021/acsomega.0c05469
  4. Angela B. Grommet, Lucia M. Lee, Rafal Klajn. Molecular Photoswitching in Confined Spaces. Accounts of Chemical Research 2020, 53 (11) , 2600-2610. https://doi.org/10.1021/acs.accounts.0c00434
  5. Julius Gemen, Johannes Ahrens, Linda J. W. Shimon, Rafal Klajn. Modulating the Optical Properties of BODIPY Dyes by Noncovalent Dimerization within a Flexible Coordination Cage. Journal of the American Chemical Society 2020, 142 (41) , 17721-17729. https://doi.org/10.1021/jacs.0c08589
  6. Cristina García-Simón, Cédric Colomban, Yarkin Aybars Çetin, Ana Gimeno, Míriam Pujals, Ernest Ubasart, Carles Fuertes-Espinosa, Karam Asad, Nikos Chronakis, Miquel Costas, Jesús Jiménez-Barbero, Ferran Feixas, Xavi Ribas. Complete Dynamic Reconstruction of C60, C70, and (C59N)2 Encapsulation into an Adaptable Supramolecular Nanocapsule. Journal of the American Chemical Society 2020, 142 (37) , 16051-16063. https://doi.org/10.1021/jacs.0c07591
  7. Martina Canton, Angela B. Grommet, Luca Pesce, Julius Gemen, Shiming Li, Yael Diskin-Posner, Alberto Credi, Giovanni M. Pavan, Joakim Andréasson, Rafal Klajn. Improving Fatigue Resistance of Dihydropyrene by Encapsulation within a Coordination Cage. Journal of the American Chemical Society 2020, 142 (34) , 14557-14565. https://doi.org/10.1021/jacs.0c06146
  8. Yan Sun, Peter J. Stang. Metallacycles, metallacages, and their aggregate/optical behavior. Aggregate 2021, 53 https://doi.org/10.1002/agt2.94
  9. Ricard López‐Coll, Rubén Álvarez‐Yebra, Ferran Feixas, Agustí Lledó. Comprehensive Characterization of the Self‐Folding Cavitand Dynamics. Chemistry – A European Journal 2021, 104 https://doi.org/10.1002/chem.202100563
  10. Robin Küng, Tobias Pausch, Dustin Rasch, Robert Göstl, Bernd M. Schmidt. Mechanochemische Freisetzung nichtkovalent gebundener Gäste aus einem mit Polymerketten dekorierten supramolekularen Käfig. Angewandte Chemie 2021, 133 (24) , 13738-13742. https://doi.org/10.1002/ange.202102383
  11. Robin Küng, Tobias Pausch, Dustin Rasch, Robert Göstl, Bernd M. Schmidt. Mechanochemical Release of Non‐Covalently Bound Guests from a Polymer‐Decorated Supramolecular Cage. Angewandte Chemie International Edition 2021, 60 (24) , 13626-13630. https://doi.org/10.1002/anie.202102383
  12. Saeed Amirjalayer. On the Molecular Mechanism of a Photo‐Responsive Phase Change Memory. Advanced Theory and Simulations 2021, 4 (5) , 2100017. https://doi.org/10.1002/adts.202100017
  13. Xianhui Tang, Dandan Chu, Wei Gong, Yong Cui, Yan Liu. Metal‐Organic Cages with Missing Linker Defects. Angewandte Chemie International Edition 2021, 60 (16) , 9099-9105. https://doi.org/10.1002/anie.202017244
  14. Xianhui Tang, Dandan Chu, Wei Gong, Yong Cui, Yan Liu. Metal‐Organic Cages with Missing Linker Defects. Angewandte Chemie 2021, 133 (16) , 9181-9187. https://doi.org/10.1002/ange.202017244
  15. Taotao Hao, Yongsheng Yang, Wenting Liang, Chunying Fan, Xin Wang, Wanhua Wu, Xiaochuan Chen, Haiyan Fu, Hua Chen, Cheng Yang. Trace mild acid-catalysed Z → E isomerization of norbornene-fused stilbene derivatives: intelligent chiral molecular photoswitches with controllable self-recovery. Chemical Science 2021, 12 (7) , 2614-2622. https://doi.org/10.1039/D0SC05213B
  16. Marieke Gerth, José Augusto Berrocal, Davide Bochicchio, Giovanni M. Pavan, Ilja K. Voets. Discordant Supramolecular Fibres Reversibly Depolymerised by Temperature and Light. Chemistry – A European Journal 2021, 27 (5) , 1829-1838. https://doi.org/10.1002/chem.202004115
  17. Dominique Matt, Jack Harrowfield. Phosphines and other P(III)‐derivatives with Cavity‐shaped Subunits: Valuable Ligands for Supramolecular Metal Catalysis, Metal Confinement and Subtle Steric Control. ChemCatChem 2021, 13 (1) , 153-168. https://doi.org/10.1002/cctc.202001242
  18. Luka Ðorđević, Lorenzo Casimiro, Nicola Demitri, Massimo Baroncini, Serena Silvi, Francesca Arcudi, Alberto Credi, Maurizio Prato. Light‐Controlled Regioselective Synthesis of Fullerene Bis‐Adducts. Angewandte Chemie 2021, 133 (1) , 317-324. https://doi.org/10.1002/ange.202009235
  19. Luka Ðorđević, Lorenzo Casimiro, Nicola Demitri, Massimo Baroncini, Serena Silvi, Francesca Arcudi, Alberto Credi, Maurizio Prato. Light‐Controlled Regioselective Synthesis of Fullerene Bis‐Adducts. Angewandte Chemie International Edition 2021, 60 (1) , 313-320. https://doi.org/10.1002/anie.202009235
  20. Zhiyao Yang, Zejiang Liu, Lihua Yuan. Recent Advances of Photoresponsive Supramolecular Switches. Asian Journal of Organic Chemistry 2021, 10 (1) , 74-90. https://doi.org/10.1002/ajoc.202000501
  21. Wei-Bin Yu, Feng-Yi Qiu, Po Sun, Hua-Tian Shi, Zhi-Feng Xin. A new supramolecular catalytic system: the self-assembly of Rh8 cage host anthracene molecules for [4 + 4] cycloaddition induced by UV irradiation. Dalton Transactions 2020, 49 (28) , 9688-9693. https://doi.org/10.1039/D0DT01978J
  • Abstract

    Figure 1

    Figure 1. Model host–guest systems. (A, top) Structure of the supramolecular cage studied herein, formed via the self-assembly of triimidazole-based donors and cis-blocked Pd acceptors. (A, bottom) Atomistic model of the supramolecular host cage, along with a schematic representation of its octahedral structure, which can be described by the axial and equatorial distances, D1 (red) and D2 (green). (B) Structural formulas and atomistic models of the guests studied herein: azobenzene (AZB), methoxylated azobenzene (M-AZB), fluorinated azobenzene (F-AZB), and arylazopyrazole (AZP) (here shown as trans isomers).

    Figure 2

    Figure 2. Conformational free-energy landscape of the empty cage. (A) Free-energy surface (FES) as a function of D1 (distance between the axial/red Pd atoms) and D2 (distance between the midpoints of opposite edges of the cage identified by the equatorial/green Pd atoms). The color scale in the FES indicates the free-energy associated with cage conformations on the D1D2 plane (scale and legend shown in B). Four representative snapshots are shown: the starting, extended configuration (top right) corresponding to the crystal structure of the cage,(62) the energetic minimum of the FES (top left), a D1-elongated structure (bottom right), and a D2-elongated structure (bottom left). Axial and equatorial Pd atoms are colored in red and green, respectively, while the connectivity scheme is colored in orange to facilitate interpretation of the structures. (B) Probability associated with all cage conformations as a function of the relative free-energy (bin width, 0.5 kcal/mol).

    Figure 3

    Figure 3. Free-energy cost of guest encapsulation. Representative equilibrium conformations (in the D1D2 plane) of the cage encapsulating different trans (left) or cis (right) guests. For each host–guest system, we report the position of the minimum-energy conformation (colored points) and the associated isolines (same colors) enclosing all conformations within 0.5 kcal/mol from the minimum of each system. The data are projected onto the FES of the empty cage (same as Figure 2), for which we also indicate the global minimum and associated 0.5 kcal/mol isoline (in white).

    Figure 4

    Figure 4. Transcis transitions of azo-switches inside and outside the cage. (A) Kinetics of transcis isomerization of the excited M-AZB (top) and F-AZB (bottom) outside the cage (in solution) reported as examples. The measured transition times, reported below the isomerization arrows, are obtained from MD simulations using an atomistic model where the CNNC dihedral potential term for the trans-azobenzene derivatives (Edih, reported in the plot as a function of the dihedral angle), is changed from the black curve (native/unperturbed state) to the blue curve (excited trans-azobenzene, S*).(50) (B) Kinetics of transcis isomerization of excited M-AZB (top) and F-AZB (bottom) switches confined inside the cage. Transition times for all the guests in the cage are reported in Table 1.

    Figure 5

    Figure 5. Thermodynamics and kinetics of trans guest binding/release. (A) Representative MetaD snapshots of the reversible binding and release of trans-M-AZB inside the cage; koff and kon denote the kinetic constants for the expulsion and encapsulation processes (the kon value inside the brackets is explicitly calculated, accounting for the guest concentration used in the model ∼11.4 mM, providing the actual rate). (B) Thermodynamic and kinetic scheme representing the expulsion and encapsulation mechanisms for trans-AZB in/out the cage as a function of the distance between the guests’ and the cage’s centers of mass (identifying IN and OUT states). (C) Thermodynamic schemes representing the expulsion/encapsulation of the trans isomers of M-AZB (green), F-AZB (cyan), and AZP (violet) guests. All ΔG differences between IN vs OUT states were computed from converged MetaD simulations, while the transition barriers were more accurately estimated from multiple infrequent MetaD runs (see the SI Methods section for further details).

    Figure 6

    Figure 6. Molecular determinants of isomerization under confinement. (A) Relationship between isomerization rate (τtrans–cis) and residence times (τoff) of the guests inside the cage. (B) Relationship between τtrans–cis and potential energy of host–guest interactions, ΔEHG. (C) Relationship between τtrans–cis and the number of contacts between the cage and the guest. (D) Relationship between τtrans–cis and the volume (V) of guest molecules (see the SI Methods section for details on guest volume estimation). (E) Switching deceleration, τtrans–cis0, as a function of the increase in guest volume (%ΔVguest), in which τ0 denotes isomerization time measured at the original volume of each guest. In plots A–E, the points correspond to guests AZB (black), M-AZB (green), F-AZB (cyan), and AZP (violet). (F) Average τtrans–cis0 as a function of the average increase in guest volume (%ΔVguest), obtained by averaging all data from plot E between systems with similar %ΔVguest. The error bars indicate the standard deviation of %ΔVguest and τtrans–cis0 values. Inset: cartoon showing the volume of encapsulated M-AZB (green) inside the cage (white). The dashed lines in all plots are the logarithmic fit of the data.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 76 other publications.

    1. 1
      Bandara, H. M. D.; Burdette, S. C. Photoisomerization in Different Classes of Azobenzene. Chem. Soc. Rev. 2012, 41, 18091825,  DOI: 10.1039/C1CS15179G
    2. 2
      Shi, Z.; Peng, P.; Strohecker, D.; Liao, Y. Long-Lived Photoacid Based upon a Photochromic Reaction. J. Am. Chem. Soc. 2011, 133, 1469914703,  DOI: 10.1021/ja203851c
    3. 3
      Klajn, R. Spiropyran-Based Dynamic Materials. Chem. Soc. Rev. 2014, 43, 148184,  DOI: 10.1039/C3CS60181A
    4. 4
      Weston, C. E.; Richardson, R. D.; Haycock, P. R.; White, A. J. P.; Fuchter, M. J. Arylazopyrazoles: Azoheteroarene Photoswitches Offering Quantitative Isomerization and Long Thermal Half-Lives. J. Am. Chem. Soc. 2014, 136, 1187811881,  DOI: 10.1021/ja505444d
    5. 5
      Wang, Y.-T.; Liu, X.-Y.; Cui, G.; Fang, W.-H.; Thiel, W. Photoisomerization of Arylazopyrazole Photoswitches: Stereospecific Excited-State Relaxation. Angew. Chem., Int. Ed. 2016, 55, 1400914013,  DOI: 10.1002/anie.201607373
    6. 6
      Fredy, J. W.; Méndez-Ardoy, A.; Kwangmettatam, S.; Bochicchio, D.; Matt, B.; Stuart, M. C. A.; Huskens, J.; Katsonis, N.; Pavan, G. M.; Kudernac, T. Molecular Photoswitches Mediating the Strain-Driven Disassembly of Supramolecular Tubules. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 1185011855,  DOI: 10.1073/pnas.1711184114
    7. 7
      Yagai, S.; Iwai, K.; Yamauchi, M.; Karatsu, T.; Kitamura, A.; Uemura, S.; Morimoto, M.; Wang, H.; Würthner, F. Photocontrol Over Self-Assembled Nanostructures of π-π Stacked Dyes Supported by the Parallel Conformer of Diarylethene. Angew. Chem., Int. Ed. 2014, 53, 26022606,  DOI: 10.1002/anie.201310773
    8. 8
      Molla, M. R.; Rangadurai, P.; Antony, L.; Swaminathan, S.; de Pablo, J. J.; Thayumanavan, S. Dynamic Actuation of Glassy Polymersomes through Isomerization of a Single Azobenzene Unit at the Block Copolymer Interface. Nat. Chem. 2018, 10, 659666,  DOI: 10.1038/s41557-018-0027-6
    9. 9
      Pianowski, Z. L. Recent Implementations of Molecular Photoswitches into Smart Materials and Biological Systems. Chem. - Eur. J. 2019, 25, 51285144,  DOI: 10.1002/chem.201805814
    10. 10
      Feringa, B. L. The Art of Building Small: From Molecular Switches to Motors (Nobel Lecture). Angew. Chem., Int. Ed. 2017, 56, 1106011078,  DOI: 10.1002/anie.201702979
    11. 11
      Sauvage, J.-P. From Chemical Topology to Molecular Machines (Nobel Lecture). Angew. Chem., Int. Ed. 2017, 56, 1108011093,  DOI: 10.1002/anie.201702992
    12. 12
      Stoddart, J. F. Mechanically Interlocked Molecules (MIMs)—Molecular Shuttles, Switches, and Machines (Nobel Lecture). Angew. Chem., Int. Ed. 2017, 56, 1109411125,  DOI: 10.1002/anie.201703216
    13. 13
      Dri, C.; Peters, M. V.; Schwarz, J.; Hecht, S.; Grill, L. Spatial Periodicity in Molecular Switching. Nat. Nanotechnol. 2008, 3, 649653,  DOI: 10.1038/nnano.2008.269
    14. 14
      Kassem, S.; Lee, A. T. L.; Leigh, D. A.; Marcos, V.; Palmer, L. I.; Pisano, S. Stereodivergent Synthesis with a Programmable Molecular Machine. Nature 2017, 549, 374378,  DOI: 10.1038/nature23677
    15. 15
      Katsonis, N.; Lubomska, M.; Pollard, M. M.; Feringa, B. L.; Rudolf, P. Synthetic Light-Activated Molecular Switches and Motors on Surfaces. Prog. Surf. Sci. 2007, 82, 407434,  DOI: 10.1016/j.progsurf.2007.03.011
    16. 16
      Broichhagen, J.; Frank, J. A.; Trauner, D. A Roadmap to Success in Photopharmacology. Acc. Chem. Res. 2015, 48, 19471960,  DOI: 10.1021/acs.accounts.5b00129
    17. 17
      Weston, C. E.; Krämer, A.; Colin, F.; Yildiz, Ö.; Baud, M. G. J.; Meyer-Almes, F.-J.; Fuchter, M. J. Toward Photopharmacological Antimicrobial Chemotherapy Using Photoswitchable Amidohydrolase Inhibitors. ACS Infect. Dis. 2017, 3, 152161,  DOI: 10.1021/acsinfecdis.6b00148
    18. 18
      Fiedler, D.; Leung, D.; Bergman, R.; Raymond, K. Selective Molecular Recognition, C-H Bond Activation, and Catalysis in Nanoscale Reaction Vessels. Acc. Chem. Res. 2005, 38, 349358,  DOI: 10.1021/ar040152p
    19. 19
      Tripp, B. C.; Smith, K.; Ferry, J. G. Carbonic Anhydrase: New Insights for an Ancient Enzyme. J. Biol. Chem. 2001, 276, 4861548618,  DOI: 10.1074/jbc.R100045200
    20. 20
      Zhang, X.; Houk, K. N. Why Enzymes Are Proficient Catalysts: Beyond the Pauling Paradigm. Acc. Chem. Res. 2005, 38, 379385,  DOI: 10.1021/ar040257s
    21. 21
      Whicher, J. R.; Dutta, S.; Hansen, D. A.; Hale, W. A.; Chemler, J. A.; Dosey, A. M.; Narayan, A. R. H.; Håkansson, K.; Sherman, D. H.; Smith, J. L.; Skiniotis, G. Structural Rearrangements of a Polyketide Synthase Module during Its Catalytic Cycle. Nature 2014, 510, 560564,  DOI: 10.1038/nature13409
    22. 22
      Grommet, A. B.; Feller, M.; Klajn, R. Chemical Reactivity under Nanoconfinement. Nat. Nanotechnol. 2020, 15, 256271,  DOI: 10.1038/s41565-020-0652-2
    23. 23
      Janiak, C.; Vieth, J. K. MOFs, MILs and More: Concepts, Properties and Applications for Porous Coordination Networks (PCNs). New J. Chem. 2010, 34, 23662388,  DOI: 10.1039/c0nj00275e
    24. 24
      Goettmann, F.; Sanchez, C. How Does Confinement Affect the Catalytic Activity of Mesoporous Materials?. J. Mater. Chem. 2007, 17, 2430,  DOI: 10.1039/B608748P
    25. 25
      Boyd, P. G.; Chidambaram, A.; García-Díez, E.; Ireland, C. P.; Daff, T. D.; Bounds, R.; Gładysiak, A.; Schouwink, P.; Moosavi, S. M.; Maroto-Valer, M. M.; Reimer, J. A.; Navarro, J. A. R.; Woo, T. K.; Garcia, S.; Stylianou, K. C.; Smit, B. Data-Driven Design of Metal–Organic Frameworks for Wet Flue Gas CO2 Capture. Nature 2019, 576, 253256,  DOI: 10.1038/s41586-019-1798-7
    26. 26
      Chu, Z.; Han, Y.; Bian, T.; De, S.; Král, P.; Klajn, R. Supramolecular Control of Azobenzene Switching on Nanoparticles. J. Am. Chem. Soc. 2019, 141, 19491960,  DOI: 10.1021/jacs.8b09638
    27. 27
      Ahrens, J.; Bian, T.; Vexler, T.; Klajn, R. Irreversible Bleaching of Donor–Acceptor Stenhouse Adducts on the Surfaces of Magnetite Nanoparticles. ChemPhotoChem. 2017, 1, 230236,  DOI: 10.1002/cptc.201700009
    28. 28
      Zdobinsky, T.; Sankar Maiti, P.; Klajn, R. Support Curvature and Conformational Freedom Control Chemical Reactivity of Immobilized Species. J. Am. Chem. Soc. 2014, 136, 27112714,  DOI: 10.1021/ja411573a
    29. 29
      Zhao, H.; Sen, S.; Udayabhaskararao, T.; Sawczyk, M.; Kučanda, K.; Manna, D.; Kundu, P. K.; Lee, J.-W.; Král, P.; Klajn, R. Reversible Trapping and Reaction Acceleration within Dynamically Self-Assembling Nanoflasks. Nat. Nanotechnol. 2016, 11, 8288,  DOI: 10.1038/nnano.2015.256
    30. 30
      Fallah-Araghi, A.; Meguellati, K.; Baret, J.-C.; Harrak, A. E.; Mangeat, T.; Karplus, M.; Ladame, S.; Marques, C. M.; Griffiths, A. D. Enhanced Chemical Synthesis at Soft Interfaces: A Universal Reaction-Adsorption Mechanism in Microcompartments. Phys. Rev. Lett. 2014, 112, 028301  DOI: 10.1103/PhysRevLett.112.028301
    31. 31
      Franco, C.; Rodríguez-San-Miguel, D.; Sorrenti, A.; Sevim, S.; Pons, R.; Platero-Prats, A. E.; Pavlovic, M.; Szilágyi, I.; Ruiz Gonzalez, M. L.; González-Calbet, J. M.; Bochicchio, D.; Pesce, L.; Pavan, G. M.; Imaz, I.; Cano-Sarabia, M.; Maspoch, D.; Pané, S.; de Mello, A. J.; Zamora, F.; Puigmartí-Luis, J. Biomimetic Synthesis of Sub-20 nm Covalent Organic Frameworks in Water. J. Am. Chem. Soc. 2020, 142, 35403547,  DOI: 10.1021/jacs.9b12389
    32. 32
      Zhang, G.; Mastalerz, M. Organic Cage Compounds – from Shape-Persistency to Function. Chem. Soc. Rev. 2014, 43, 19341947,  DOI: 10.1039/C3CS60358J
    33. 33
      Kang, J.; Rebek, J. Acceleration of a Diels–Alder Reaction by a Self-Assembled Molecular Capsule. Nature 1997, 385, 5052,  DOI: 10.1038/385050a0
    34. 34
      Yoshizawa, M.; Klosterman, J. K.; Fujita, M. Functional Molecular Flasks: New Properties and Reactions within Discrete, Self-Assembled Hosts. Angew. Chem., Int. Ed. 2009, 48, 34183438,  DOI: 10.1002/anie.200805340
    35. 35
      Roy, B.; Ghosh, A. K.; Srivastava, S.; D’Silva, P.; Mukherjee, P. S. A Pd8 Tetrafacial Molecular Barrel as Carrier for Water Insoluble Fluorophore. J. Am. Chem. Soc. 2015, 137, 1191611919,  DOI: 10.1021/jacs.5b08008
    36. 36
      Wang, K.; Cai, X.; Yao, W.; Tang, D.; Kataria, R.; Ashbaugh, H. S.; Byers, L. D.; Gibb, B. C. Electrostatic Control of Macrocyclization Reactions within Nanospaces. J. Am. Chem. Soc. 2019, 141, 67406747,  DOI: 10.1021/jacs.9b02287
    37. 37
      Cook, T. R.; Stang, P. J. Recent Developments in the Preparation and Chemistry of Metallacycles and Metallacages via Coordination. Chem. Rev. 2015, 115, 70017045,  DOI: 10.1021/cr5005666
    38. 38
      Sepehrpour, H.; Fu, W.; Sun, Y.; Stang, P. J. Biomedically Relevant Self-Assembled Metallacycles and Metallacages. J. Am. Chem. Soc. 2019, 141, 1400514020,  DOI: 10.1021/jacs.9b06222
    39. 39
      Liu, M.; Zhang, L.; Little, M. A.; Kapil, V.; Ceriotti, M.; Yang, S.; Ding, L.; Holden, D. L.; Balderas-Xicohténcatl, R.; He, D.; Clowes, R.; Chong, S. Y.; Schütz, G.; Chen, L.; Hirscher, M.; Cooper, A. I. Barely Porous Organic Cages for Hydrogen Isotope Separation. Science 2019, 366, 613620,  DOI: 10.1126/science.aax7427
    40. 40
      Merget, S.; Catti, L.; Piccini, G.; Tiefenbacher, K. Requirements for Terpene Cyclizations inside the Supramolecular Resorcinarene Capsule: Bound Water and Its Protonation Determine the Catalytic Activity. J. Am. Chem. Soc. 2020, 142, 44004410,  DOI: 10.1021/jacs.9b13239
    41. 41
      Yoshizawa, M.; Tamura, M.; Fujita, M. Diels-Alder in Aqueous Molecular Hosts: Unusual Regioselectivity and Efficient Catalysis. Science 2006, 312, 251254,  DOI: 10.1126/science.1124985
    42. 42
      Ueda, Y.; Ito, H.; Fujita, D.; Fujita, M. Permeable Self-Assembled Molecular Containers for Catalyst Isolation Enabling Two-Step Cascade Reactions. J. Am. Chem. Soc. 2017, 139, 60906093,  DOI: 10.1021/jacs.7b02745
    43. 43
      Maestri, M.; Iglesia, E. First-Principles Theoretical Assessment of Catalysis by Confinement: NO–O2 Reactions within Voids of Molecular Dimensions in Siliceous Crystalline Frameworks. Phys. Chem. Chem. Phys. 2018, 20, 1572515735,  DOI: 10.1039/C8CP01615A
    44. 44
      Mal, P.; Breiner, B.; Rissanen, K.; Nitschke, J. R. White Phosphorus Is Air-Stable Within a Self-Assembled Tetrahedral Capsule. Science 2009, 324, 1697,  DOI: 10.1126/science.1175313
    45. 45
      Yamashina, M.; Sei, Y.; Akita, M.; Yoshizawa, M. Safe Storage of Radical Initiators within a Polyaromatic Nanocapsule. Nat. Commun. 2014, 5, 4662,  DOI: 10.1038/ncomms5662
    46. 46
      Galan, A.; Ballester, P. Stabilization of Reactive Species by Supramolecular Encapsulation. Chem. Soc. Rev. 2016, 45, 17201737,  DOI: 10.1039/C5CS00861A
    47. 47
      Qiu, Y.; Antony, L. W.; Torkelson, J. M.; de Pablo, J. J.; Ediger, M. D. Tenfold Increase in the Photostability of an Azobenzene Guest in Vapor-Deposited Glass Mixtures. J. Chem. Phys. 2018, 149, 204503,  DOI: 10.1063/1.5052003
    48. 48
      Qiu, Y.; Antony, L. W.; de Pablo, J. J.; Ediger, M. D. Photostability Can Be Significantly Modulated by Molecular Packing in Glasses. J. Am. Chem. Soc. 2016, 138, 1128211289,  DOI: 10.1021/jacs.6b06372
    49. 49
      Fregoni, J.; Granucci, G.; Persico, M.; Corni, S. Strong Coupling with Light Enhances the Photoisomerization Quantum Yield of Azobenzene. Chem. 2020, 6, 250265,  DOI: 10.1016/j.chempr.2019.11.001
    50. 50
      Bochicchio, D.; Kwangmettatam, S.; Kudernac, T.; Pavan, G. M. How Defects Control the Out-of-Equilibrium Dissipative Evolution of a Supramolecular Tubule. ACS Nano 2019, 13, 43224334,  DOI: 10.1021/acsnano.8b09523
    51. 51
      Kusukawa, T.; Fujita, M. Ship-in-a-Bottle” Formation of Stable Hydrophobic Dimers of Cis-Azobenzene and -Stilbene Derivatives in a Self-Assembled Coordination Nanocage. J. Am. Chem. Soc. 1999, 121, 13971398,  DOI: 10.1021/ja9837295
    52. 52
      Cantatore, V.; Granucci, G.; Rousseau, G.; Padula, G.; Persico, M. Photoisomerization of Self-Assembled Monolayers of Azobiphenyls: Simulations Highlight the Role of Packing and Defects. J. Phys. Chem. Lett. 2016, 7, 40274031,  DOI: 10.1021/acs.jpclett.6b02018
    53. 53
      Clever, G. H.; Tashiro, S.; Shionoya, M. Light-Triggered Crystallization of a Molecular Host-Guest Complex. J. Am. Chem. Soc. 2010, 132, 99739975,  DOI: 10.1021/ja103620z
    54. 54
      Dube, H.; Ajami, D.; Rebek, J. Photochemical Control of Reversible Encapsulation. Angew. Chem., Int. Ed. 2010, 49, 31923195,  DOI: 10.1002/anie.201000876
    55. 55
      Mohan Raj, A.; Raymo, F. M.; Ramamurthy, V. Reversible Disassembly–Assembly of Octa Acid–Guest Capsule in Water Triggered by a Photochromic Process. Org. Lett. 2016, 18, 15661569,  DOI: 10.1021/acs.orglett.6b00405
    56. 56
      Yang, Y.; Hughes, R. P.; Aprahamian, I. Visible Light Switching of a BF2-Coordinated Azo Compound. J. Am. Chem. Soc. 2012, 134, 1522115224,  DOI: 10.1021/ja306030d
    57. 57
      Helmy, S.; Leibfarth, F. A.; Oh, S.; Poelma, J. E.; Hawker, C. J.; Read de Alaniz, J. Photoswitching Using Visible Light: A New Class of Organic Photochromic Molecules. J. Am. Chem. Soc. 2014, 136, 81698172,  DOI: 10.1021/ja503016b
    58. 58
      Zhang, D.; Ronson, T. K.; Mosquera, J.; Martinez, A.; Guy, L.; Nitschke, J. R. Anion Binding in Water Drives Structural Adaptation in an Azaphosphatrane-Functionalized FeII4L4 Tetrahedron. J. Am. Chem. Soc. 2017, 139, 65746577,  DOI: 10.1021/jacs.7b02950
    59. 59
      Rizzuto, F. J.; Nitschke, J. R. Stereochemical Plasticity Modulates Cooperative Binding in a CoII12L6 Cuboctahedron. Nat. Chem. 2017, 9, 903908,  DOI: 10.1038/nchem.2758
    60. 60
      Mondal, P.; Sarkar, S.; Rath, S. P. Cyclic Bis-Porphyrin-Based Flexible Molecular Containers: Controlling Guest Arrangements and Supramolecular Catalysis by Tuning Cavity Size. Chem. - Eur. J. 2017, 23, 70937103,  DOI: 10.1002/chem.201700577
    61. 61
      Samanta, D.; Mukherjee, S.; Patil, Y. P.; Mukherjee, P. S. Self-Assembled Pd6 Open Cage with Triimidazole Walls and the Use of Its Confined Nanospace for Catalytic Knoevenagel- and Diels–Alder Reactions in Aqueous Medium. Chem. - Eur. J. 2012, 18, 1232212329,  DOI: 10.1002/chem.201201679
    62. 62
      Samanta, D.; Gemen, J.; Chu, Z.; Diskin-Posner, Y.; Shimon, L. J. W.; Klajn, R. Reversible Photoswitching of Encapsulated Azobenzenes in Water. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 93799384,  DOI: 10.1073/pnas.1712787115
    63. 63
      Samanta, D.; Galaktionova, D.; Gemen, J.; Shimon, L. J. W.; Diskin-Posner, Y.; Avram, L.; Král, P.; Klajn, R. Reversible Chromism of Spiropyran in the Cavity of a Flexible Coordination Cage. Nat. Commun. 2018, 9, 641,  DOI: 10.1038/s41467-017-02715-6
    64. 64
      Hanopolskyi, A. I.; De, S.; Białek, M. J.; Diskin-Posner, Y.; Avram, L.; Feller, M.; Klajn, R. Reversible Switching of Arylazopyrazole within a Metal–Organic Cage. Beilstein J. Org. Chem. 2019, 15, 23982407,  DOI: 10.3762/bjoc.15.232
    65. 65
      Böckmann, M.; Peter, C.; Site, L. D.; Doltsinis, N. L.; Kremer, K.; Marx, D. Atomistic Force Field for Azobenzene Compounds Adapted for QM/MM Simulations with Applications to Liquids and Liquid Crystals. J. Chem. Theory Comput. 2007, 3, 17891802,  DOI: 10.1021/ct7000733
    66. 66
      Peter, C.; Site, L. D.; Kremer, K. Classical Simulations from the Atomistic to the Mesoscale and Back: Coarse Graining an Azobenzene Liquid Crystal. Soft Matter 2008, 4, 859869,  DOI: 10.1039/b717324e
    67. 67
      Ilnytskyi, J. M.; Slyusarchuk, A.; Saphiannikova, M. Photocontrollable Self-Assembly of Azobenzene-Decorated Nanoparticles in Bulk: Computer Simulation Study. Macromolecules 2016, 49, 92729282,  DOI: 10.1021/acs.macromol.6b01871
    68. 68
      Osella, S.; Minoia, A.; Beljonne, D. Combined Molecular Dynamics and Density Functional Theory Study of Azobenzene–Graphene Interfaces. J. Phys. Chem. C 2016, 120, 66516658,  DOI: 10.1021/acs.jpcc.6b00393
    69. 69
      Döbbelin, M.; Ciesielski, A.; Haar, S.; Osella, S.; Bruna, M.; Minoia, A.; Grisanti, L.; Mosciatti, T.; Richard, F.; Prasetyanto, E. A.; De Cola, L.; Palermo, V.; Mazzaro, R.; Morandi, V.; Lazzaroni, R.; Ferrari, A. C.; Beljonne, D.; Samorì, P. Light-Enhanced Liquid-Phase Exfoliation and Current Photoswitching in Graphene–Azobenzene Composites. Nat. Commun. 2016, 7, 11090,  DOI: 10.1038/ncomms11090
    70. 70
      Laio, A.; Parrinello, M. Escaping Free-Energy Minima. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 1256212566,  DOI: 10.1073/pnas.202427399
    71. 71
      Pederzoli, M.; Pittner, J.; Barbatti, M.; Lischka, H. Nonadiabatic Molecular Dynamics Study of the CisTrans Photoisomerization of Azobenzene Excited to the S1 State. J. Phys. Chem. A 2011, 115, 1113611143,  DOI: 10.1021/jp2013094
    72. 72
      Tiago, M. L.; Ismail-Beigi, S.; Louie, S. G. Photoisomerization of Azobenzene from First-Principles Constrained Density-Functional Calculations. J. Chem. Phys. 2005, 122, 094311  DOI: 10.1063/1.1861873
    73. 73
      Bochicchio, D.; Salvalaglio, M.; Pavan, G. M. Into the Dynamics of a Supramolecular Polymer at Submolecular Resolution. Nat. Commun. 2017, 8, 147,  DOI: 10.1038/s41467-017-00189-0
    74. 74
      Tiwary, P.; Parrinello, M. From Metadynamics to Dynamics. Phys. Rev. Lett. 2013, 111, 230602,  DOI: 10.1103/PhysRevLett.111.230602
    75. 75
      Pace, G.; Ferri, V.; Grave, C.; Elbing, M.; von Hänisch, C.; Zharnikov, M.; Mayor, M.; Rampi, M. A.; Samorì, P. Cooperative Light-Induced Molecular Movements of Highly Ordered Azobenzene Self-Assembled Monolayers. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 9937,  DOI: 10.1073/pnas.0703748104
    76. 76
      Titov, E.; Granucci, G.; Götze, J. P.; Persico, M.; Saalfrank, P. Dynamics of Azobenzene Dimer Photoisomerization: Electronic and Steric Effects. J. Phys. Chem. Lett. 2016, 7, 35913596,  DOI: 10.1021/acs.jpclett.6b01401
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.0c03444.

    • Details on the creation and parametrization of the molecular systems, simulation setup, and analysis of molecular dynamics and metadynamics simulations; additional data and figures from the simulations (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect

This website uses cookies to improve your user experience. By continuing to use the site, you are accepting our use of cookies. Read the ACS privacy policy.

CONTINUE