You are on page 1of 6

Letter

Cite This: J. Phys. Chem. Lett. 2018, 9, 1448−1453 pubs.acs.org/JPCL

Sensitized Two-Photon Activation of Coumarin Photocages


Christopher A. Hammer,†,§ Konstantin Falahati,†,§ Andreas Jakob,‡,§ Robin Klimek,‡ Irene Burghardt,†
Alexander Heckel,*,‡ and Josef Wachtveitl*,†

Institute of Physical and Theoretical Chemistry and ‡Institute of Organic Chemistry and Chemical Biology, Goethe University
Frankfurt am Main, Max-von-Laue-Straße 7, 60438 Frankfurt am Main, Germany
*
S Supporting Information

ABSTRACT: Here we report the design of a new coumarin-based


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

photolabile protecting group with enhanced two-photon absorption. Two-


photon excited fluorescence (TPEF), color-tuned ultrafast transient
absorption spectroscopy and infrared (IR) measurements are employed
to photochemically characterize the newly designed ATTO 390-DEACM-
Downloaded via NATL TAIWAN UNIV on July 3, 2022 at 15:17:05 (UTC).

cargo triad. Increased two-photon cross-section values of the novel cage in


comparison to the widely used protecting group DEACM ([7-
(diethylamino)coumarin-4-yl]methyl) are extracted from TPEF experi-
ments. Femtosecond pump−probe experiments reveal a fast intramolecular
charge transfer, a finding that is confirmed by quantum chemical
calculations. Uncaging of glutamate is monitored in IR measurements by
photodecarboxylation of the carbamate linker between the photolabile
protecting group and the glutamate, showing the full functionality of the
novel two-photon activatable photocage.

I n many areas of material and life sciences, massive interest to


control biological and chemical systems have emerged in the
past few years. Light-induced triggering in such systems is a
window” (650−950 nm) are accessible, which are suitable for
deeper penetration into blood-perfused tissue with lower
phototoxicity.28,29 By now, the field shows considerable
promising approach to realize highly specific control promise, but only few potent two-photon activatable photo-
strategies.1−3 The ability to mask the biological activity of a labile protecting groups are available to date, and the rational
compound by attaching a photolabile protecting group design or computational prediction of two-photon properties is
(“photocage” or “caging group”) and set it to its active or still a formidable challenge. Hence, the present study addresses
inactive state by irradiating the photocage with light has been a novel molecular design strategy aimed at achieving an
exploited for biological applications since the end of the enhanced two-photon absorption of the DEACM cage. Our
1970s.4,5 Since then, many interesting examples have emerged concept is based on the observation that fluorophores with a
using photocages or photoswitches and have demonstrated, for large two-photon cross-section are readily available, whereas the
example, light control of gene expression,6,7 aptamer function,8 majority of the known photocages have not yet reached a
protein function,9 lipids10,11 or signaling molecules.12 In the comparable two-photon cross-section range.30 Hence, we
small molecule domain, the term “photopharmacology” was explore the idea to use suitable two-photon fluorophores as
coined13,1 and, for example, the light-control of the narcotic antennae to operate a photocage in an intramolecular dyad. As
propofol and its application to living tadpoles was demon- the antenna fluorophore we chose ATTO 390 (labeled I in
strated in a spectacular study.14 Figure 1A) due to its significant two-photon cross section of 14
Among the known photocages the photolabile protecting GM. The fluorophore was connected to a DEACM core via an
group DEACM exhibits outstanding properties, e.g., water alkyne-containing linker (labeled II in Figure 1A). As model
solubility, absorption in the visible part of the spectrum and the cargo to be released from this intramolecular dyad we chose
possibility to be introduced into nucleic acids or DNA strands, glutamate (labeled III in Figure 1A) for its potential use as
which explains its widespread utilization.15−20 Furthermore, the light-triggerable neurotransmitter.31,32 For the synthesis of this
uncaging mechanism of (coumarin-4-yl)methyl derivatives is compound, we refer to the SI.
well-studied and understood and therefore DEACM is the In view of working toward rational design strategies adapted
photocage of choice.19,21−23 Nonetheless, controlled triggering to supramolecular structures like I+II, the present detailed
with high spatial and temporal resolution represents a further molecular-level study combines time-resolved spectroscopy
obstacle. Techniques based on two-photon absorption are able with computational analysis. Besides examining the efficiency
to respond to this challenge, since they enable biological
applications with high 3D resolution,18,24−27 where conven- Received: December 20, 2017
tional one-photon techniques provide only two-dimensional Accepted: March 2, 2018
control. Moreover, wavelength ranges in the “phototherapeutic Published: March 2, 2018

© 2018 American Chemical Society 1448 DOI: 10.1021/acs.jpclett.7b03364


J. Phys. Chem. Lett. 2018, 9, 1448−1453
The Journal of Physical Chemistry Letters Letter

absorption bands of I and II. Nevertheless, the shift of bands of


the isolated chromophores indicates electronic interaction
within I+II. For two-photon cross sections, it is clearly
recognizable that the values of DEACM are the lowest of all
three compounds. On the other hand, I exhibits comparatively
large GM values, underlining that it is a good candidate for two-
photon sensitization. I+II exhibits at all excitation wavelengths
a higher two-photon absorption cross-section than II and even
higher values than I in the range of 820−870 nm. At this point,
it should be emphasized that TPEF is an indirect method of
measuring the two-photon absorption cross-section, which is
strongly coupled to the two-photon fluorescence quantum yield
(ϕ2F). In a first approximation, the one-photon excited
fluorescence is similar to two-photon excited fluorescence.
The one-photon excited fluorescence quantum yields (ϕ1F) of
I, II and I+II were determined to be 1, 0.85 and 0.7,
respectively (see SI for further information). By taking the ϕ1F
into account, we assume that the two-photon absorption of I
+II is even higher, an important prerequisite for two-photon
activation of the photocage.
Ultrafast UV/vis pump−probe transient absorption measure-
ments have been performed to study the photophysics of I+II.
We examined the ultrafast dynamics in a wavelength dependent
fashion by photoexciting the sample at 365 nm, 388 and 475
nm. The control of the two-photon excitation pulse diameter in
pump−probe experiments is quite challenging. Since the
dynamics of I+II should be independent of the excitation
mechanism, the photocage is excited with one photon. The
Figure 1. (A) Chemical structure and labeling of the triad. (B) resulting transient absorption spectra are depicted in Figure 2.
Absorption of ATTO 390 (I), the dyad consisting of ATTO 390 and On the basis of the absorption spectra in Figure 1B, the
DEACM with a propargyl linker in 3 position (I+II) and the latter chosen excitation wavelength of 365 nm addresses predom-
alone (II); the black triangles indicate the excitation wavelengths used inantly I, whereas an exclusive photoexcitation of II at 475 nm
in the transient absorption measurements (see Figure 2) and two-
photon absorption cross sections in values of GM for I (●), I+II (Δ),
is expected and therefore should only display the photo-
and II (□) determined with the method of two-photon excited dynamics of II. Photoexcitation at 388 nm presumably displays
fluorescence and referenced to the values of Fluorescein. For more photodynamics of both. Figure 2C shows the transient
information, see SI. absorption spectrum after photoexcitation at 475 nm, where a
negative signal (shown in blue) centered at 450 nm is visible,
which represents the ground state bleach of II. The second
of the I+II system, we aim to gain a detailed understanding of negative signal at around 500−600 nm is assigned to stimulated
the intramolecular interactions that determine the dyad’s emission (SE). The negative signal of the SE and a positive
performance. As will be discussed below, energy transfer, signal (shown in red) related to the excited state absorption
charge transfer, and electronic delocalization will be found to (ESA) of II (390−440 nm) decay with a time constant of 2.6 ps
work together to create a robust and performant supra- (Figure 2C, τ1). With this time constant, a third negative signal
molecular assembly. centered at 410 nm appears, which is assigned to the ground
Two-photon absorption cross sections are determined via the state bleach of I. Since an intramolecular energy transfer (IET)
method of two-photon excited fluorescence (TPEF).33,34 To to higher energies and as well a direct excitation of I (Figure
validate our measurements, we used three reference com- 1B) are unlikely, we deduce that an intramolecular charge
pounds (Rhodamine B, Fluorescein, and Coumarin 307), transfer (ICT) from II to I must occur. The decay of the SE
(referenced to each other) and detected broadband emission to with τ1 indicates that the molecule is no longer in the excited
diminish perturbations (see SI). state, although the ground state is not recovered since the GSB
Figure 1B shows the one-photon absorption and the two- at 450 nm is still present. The subsequent charge recombina-
photon absorption cross sections in units of GM for ATTO 390 tion leads to the decay of all transient absorption signals with a
I, the DEACM-derivative II and our newly designed two- time constant of 130 ps (Figure 2C; τ2). The transient
photon sensitized photocage I+II for excitation wavelengths absorption spectrum recorded after photoexcitation at 388 nm
ranging from 740−900 nm. Spectra of the reference depicted in Figure 2B also shows three negative signals. Similar
compounds and power-dependent measurements are provided to the excitation at 475 nm, the negative signal between 500
in the SI. The absorption band of I (black line) exhibits a and 600 nm is assigned to the SE, which decays faster than the
maximum at 390 nm, while II (green) has a maximum at GSBs around 400−475 nm, indicating an ICT as described
around 430 nm. The DEACM-derivative II alone shows a red- above. However, a long time constant τ4 is necessary to
shifted absorbance compared to DEACM due to its propargylic describe residual SE and long-wavelength GSB signals, while no
substitution in the 3-position. contribution of the short-wavelength GSB is visible at late delay
The absorption band of the dyad I+II (red) reveals a times. In the transient absorption spectrum recorded after
maximum at 400 nm and is most likely composed of the photoexcitation at 365 nm where mainly I should absorb
1449 DOI: 10.1021/acs.jpclett.7b03364
J. Phys. Chem. Lett. 2018, 9, 1448−1453
The Journal of Physical Chemistry Letters Letter

Figure 2. Transient absorption spectra of I+II excited at (A) λexc = 365 nm, (B) λexc = 388 nm, and (C) λexc = 475 nm. The gray bar indicates that
data in this range were excluded from fitting, due to the photoexcitation at 475 nm. In the lower panel, the corresponding decay associated spectra
are shown.

(Figure 2A) two major negative bands centered at 400 and 500 and energy transfer (I to II). To validate our findings in the
nm and a weak negative band at approximately 450 nm are experimental section, we performed quantum chemical
present. We assign the first negative signal (400 nm) to the calculations, as now detailed.
short-wavelength ground state bleach of I. The weak negative Ground state geometry optimization of the model system at
signal (450 nm) belongs to the long-wavelength GSB of II. the ωB97XD/SVP35,36 DFT level of theory using the
Negative signals ≥500 nm are assigned to the SE. Three Gaussian09 program package37 reveals a π-stacked aggregation
positive signals (shown in red) in the very blue and red part of of the hydrophobic moieties (see Figure 3). The relative
the spectrum and centered at 475 nm are assigned to ESAs. energetics indicate selectively high stability for the stacked
The long-wavelength GSB assigned to II is directly present, conformer both in gas phase as well as solvation model
which is most probably due to residual direct photoexcitation of embedding (see SI for further thermochemical details). A linear
II. An ultrafast energy transfer is ruled out, since the short- and hence unfolded conformation was found to be roughly 25
wavelength GSB of I is not decreasing on this time scale. kcal/mol higher in energy than the stacked conformer. Note
Between 0 and 100 ps, an increase of the GSBs and a decrease that the stacked conformation further benefits energetically
of the SE is observed. Intriguingly, the ESA at 650 nm remains from a stabilizing intramolecular hydrogen bond as indicated in
constant, implying complex dynamics in the excited state red in panel A.
described by τ1 and τ2, a relaxation to the ground state is not Subsequent excited state analysis was performed by means of
observed. At delay times of few hundred picoseconds to TD-DFT calculations in gas phase and using a solvation model.
nanoseconds, the long-wavelength GSB persists to a certain The bright states (with excitation energies at 3.6 eV (343 nm)
extent, while the short-wavelength GSB completely decreases, and 4.1 eV (306 nm), respectively) feature non-negligible ICT
indicating an IET from I to II (Figure 2A, compare DAS of τ3 character at the Franck−Condon geometry as can be inferred
and τ4). Photoexcitation at 475 nm indicates a CT from II to I from the Supporting Material, thus highlighting a close
(Figure 2C). However, the long-lived SE and the decay of the electronic entanglement of the ATTO 390 (I) and the
short-wavelength GSB is a compelling fact against an ICT from DEACM (II) subunits. The S2 state at 3.9 eV (320 nm),
II to I after the IET following photoexcitation at 365 nm. This however, features a dominant ICT character, which is reflected
feature can be also explained by direct photoexcitation of II. by an oscillator strength reduction of 1 order of magnitude.
However, Figure 2C displays that direct photoexcitation of II The mixed character of electronic excitation is indicated by the
leads to a fast recovery of the ground state. Structural changes orbital transitions in the lower panel of Figure 3. The stacked
upon photoexcitation at 365 nm could explain our observa- structure of the photocage thus prohibits exclusive local
tions. As a result of conformational changes, a charge transfer excitation of either the I or the II moiety of the molecule
from II to I after an IET from I to II is not possible anymore. effectively, a fact that matches our experimental findings
To summarize, transient absorption measurements disclose displayed in Figure 2. The picture is somewhat altered in the
the charge transfer character of photoexcited I+II. Spectral case of the (energetically unfavorable) unfolded model system:
signatures are tentatively assigned to charge transfer (II to I), TD-DFT calculations reveal rather local excitation patterns on
1450 DOI: 10.1021/acs.jpclett.7b03364
J. Phys. Chem. Lett. 2018, 9, 1448−1453
The Journal of Physical Chemistry Letters Letter

Figure 4A shows that the absorption of I+II+III changes by


continuous illumination. It can be clearly seen that the
absorbance in the wavelength range between 390 and 420
nm increases, while the shoulder in the long wavelength-range
at around 475 nm of sample I+II+III merges into the blue
region. Figure 4B shows the difference absorption spectra in the
UV/vis with an isosbestic point underlining the formation of a
photoproduct. To clarify the findings in the UV/vis region,
measurements in the IR were additionally performed. FTIR-
measurements reveal the formation of dissolved carbon dioxide
(2337 cm−1) resulting from the decomposition of the
carbamate (see Figure 4D). Moreover, the band of the carbonyl
stretch mode of the carbamate linker (1722 cm−1) decreases on
the same time scale as the formation of CO2. The release of
CO2 by uncaging a product linked via a carbamate has been
shown in the past and monitors in this case the uncaging of
glutamic acid.19,38 Accordingly, we assign the absorption change
observed in the stationary UV/vis measurements (Figure 4A)
to uncaging of glutamate indicated by the isosbestic point and
determined an uncaging quantum yield of 1.5% (IR and UV/vis
measurements on the reference compound II+III reveal an
uncaging quantum yield, which is 1 order of magnitude lower
than that of I+II+III; for further information, see the SI).
In summary, we designed a photocage with a DEACM
scaffold with enhanced two-photon absorption compared to
regular DEACM. TPEF experiments reveal larger two-photon
absorption cross sections of I+II than DEACM but not as large
as ATTO 390. Transient absorption measurements disclose the
charge transfer character of photoexcited I+II due to π-stacking
confirmed by quantum chemical calculations. Charge transfer is
detected from II to ATTO 390 (I), while most likely an energy
transfer from I to II was observed. From stationary absorption
measurements ϕunc = 1.5% was determined. FTIR-measure-
ments reveal photodecarboxylation and hence uncaging of
glutamate. We present a fully functional photocage with
enhanced two-photon absorption. This work reports the
successful intramolecular attachment of a two-photon-sensitiz-
ing moiety, resulting in a molecular triad consisting of light
harvesting unit, photocage, and effector and should help to
Figure 3. (A) Lewis structure of the calculated model system with develop future design strategies for even better two-photon
indication of ATTO 390 (I) and DEACM moieties (II). Intra- absorbing photocages.


molecular hydrogen bond highlighted in red. (B) Main orbital
transitions resembling local (blue) and CT type (red) excitation EXPERIMENTAL METHODS
character. Orbitals obtained in energetic order at the TD-ωB97XD/
SVP level of theory. The syntheses of I+II, II+III and I+II+III are described in
detail in the SI.
TPEF experiments were carried out using a tunable Ti-
either subunits of the molecule, indicating that the degree of sapphire laser (Tsunami, Spectra-Physics) producing 150 fs
ICT contribution may be tuned via a potentially relevant laser pulses with a repetition rate of 80 MHz. The fluorescence
mechanistic unstacking after initial excitation. However, since after two-photon excitation was coupled into a spectrograph
such a nuclear rearrangement upon unfolding corresponds to a (SpectraPro 300i, Acton Research Corporation) with a CCD-
considerable amount of correlated geometric motion, the camera (EEV 400_1340F, Roper Scientific).
dynamics presumably does not occur on the ultrafast time scale. Transient UV/Vis-Pump-Vis-Probe Experiments. Measure-
Nevertheless, as competing IET and ICT processes might be ments were performed using a CLARK-MXR CPA 2110 with
inferred at least from the longer time traces of Figure 2 (100 ps a central wavelength of 775 nm (pulse length 150 fs, repetition
regime), the interplay of π-stacked and unfolded conformations rate 1 kHz). The excitation pulses with a central wavelength of
can be of importance regarding further elucidation of 365 nm were generated by sum frequency of 690 nm with the
photochemical scenarios for this class of compounds. fundamental laser beam. Excitation pulses with a central
To demonstrate the functionality of the newly designed wavelength of 475 nm were generated by a home-built two
photocage, we introduced a glutamate (III) into I+II linked via stage noncollinear optical parametric amplifier (NOPA).
a carbamate to the coumarin moiety yielding compound I+II Broadband probe pulses were produced by guiding the laser
+III (see SI). We continuously irradiated the sample I+II+III fundamental trough a sapphire crystal and were split for
with an LED with a central wavelength of 365 nm and referenced measurements and subsequently recorded via
performed UV/vis and FTIR measurements (see Figure 4). photodiode arrays (Hamamatsu). For further information
1451 DOI: 10.1021/acs.jpclett.7b03364
J. Phys. Chem. Lett. 2018, 9, 1448−1453
The Journal of Physical Chemistry Letters Letter

Figure 4. Continuous illumination (λexc = 365 nm) of I+II+III monitored via (A) UV/vis spectroscopy and (E) FTIR spectroscopy (λexc = 365 nm).
(B) Absorption difference spectra of the UV/vis measurements. (C) Transient at 421 nm with a linear fit to determine the uncaging quantum yield.
The formation of dissolved CO2 is shown in panel D, and the corresponding decrease of the stretch mode of the C−O of the carbamate is depicted
in panel F.

regarding the experimental setup see ref 39. Time-resolved data Notes
obtained from transient absorption measurements were The authors declare no competing financial interest.


analyzed with OPTIMUS, where a global lifetime analysis
(GLA) was used to fit the data with a set of exponential decay ACKNOWLEDGMENTS
functions.40
Stationary Absorption Measurements. UV/vis spectra were This work was supported by the Deutsche Forschungsgemein-
recorded with a Specord S600 (Analytik Jena), while FTIR schaft (DFG) by means of the research training group “CLiC”
measurements were carried out with a VERTEX 80 (Bruker, (GRK 1986, Complex scenarios of light-control). We thank
Ettlingen). For the irradiation of the samples, a Thorlabs LED Strahinja Lucic for help with the stationary spectroscopic
characterization.


with a central wavelength of 365 nm was used.


*
ASSOCIATED CONTENT
S Supporting Information
REFERENCES
(1) Lerch, M. M.; Hansen, M. J.; van Dam, G. M.; Szymanski, W.;
Feringa, B. L. Emerging Targets in Photopharmacology. Angew. Chem.,
The Supporting Information is available free of charge on the Int. Ed. 2016, 55, 10978−10999.
ACS Publications website at DOI: 10.1021/acs.jpclett.7b03364. (2) Klán, P.; Šolomek, T.; Bochet, C. G.; Blanc, A.; Givens, R.;
Chemical synthesis; one-photon absorption, two-photon Rubina, M.; Popik, V.; Kostikov, A.; Wirz, J. Photoremovable
excited fluorescence, and wavelength-dependent TPEF Protecting Groups in Chemistry and Biology: Reaction Mechanisms
spectra of Rhodamine B, Fluorescein, and Coumarin and Efficacy. Chem. Rev. 2013, 113, 119−191.
307; power-dependent fluorescence spectra of Rhod- (3) Brieke, C.; Rohrbach, F.; Gottschalk, A.; Mayer, G.; Heckel, A.
Light-Controlled Tools. Angew. Chem., Int. Ed. 2012, 51, 8446−8476.
amine B; fit of transient absorption spectrum after (4) Engels, J.; Schlaeger, E. J. Synthesis, Structure, and Reactivity of
photoexcitation at 475 nm; calculations of uncaging and Adenosine Cyclic 3′,5′-Phosphate Benzyl Triesters. J. Med. Chem.
fluorescence quantum yield; control experiments with II 1977, 20, 907−911.
+III; quantum chemical data concerning thermochemis- (5) Kaplan, J. H.; Forbush, B., III; Hoffman, J. F. Rapid Photolytic
try and excited state constitution (PDF) Release of Adenosine 5′-triphosphate from a Protected Analogue:


Utilization by the Na:K Pump of Human Red Blood Cell Ghosts.
Biochemistry 1978, 17, 1929−1935.
AUTHOR INFORMATION (6) Lucas, T.; Schäfer, F.; Müller, P.; Eming, S. A.; Heckel, A.;
Corresponding Authors Dimmeler, S. Light-Inducible AntimiR-92a as a Therapeutic Strategy
*E-mail: heckel@uni-frankfurt.de (A.H.). to Promote Skin Repair in Healing-Impaired Diabetic Mice. Nat.
*E-mail: wveitl@theochem.uni-frankfurt.de (J.W.). Commun. 2017, 8, 15162.
(7) Ji, Y.; Yang, J.; Wu, L.; Yu, L.; Tang, X. Gene Regulation
ORCID Photochemical Regulation of Gene Expression Using Caged siRNAs
Irene Burghardt: 0000-0002-9727-9049 with Single Terminal Vitamin E Modification. Angew. Chem., Int. Ed.
Alexander Heckel: 0000-0003-3541-4548 2016, 55, 2152−2156.
Josef Wachtveitl: 0000-0002-8496-8240 (8) Rohrbach, F.; Schäfer, F.; Fichte, M. A. H.; Pfeiffer, F.; Müller, J.;
Pötzsch, B.; Heckel, A.; Mayer, G. Aptamer-Guided Caging for
Author Contributions Selective Masking of Protein Domains. Angew. Chem., Int. Ed. 2013,
§
Authors contributed equally 52, 11912−11915.

1452 DOI: 10.1021/acs.jpclett.7b03364


J. Phys. Chem. Lett. 2018, 9, 1448−1453
The Journal of Physical Chemistry Letters Letter

(9) Zhou, W.; Deiters, A. Conditional Control of CRISPR/Cas9 Protecting Groups in Chemistry and Biology: Reaction Mechanisms
Function. Angew. Chem., Int. Ed. 2016, 55, 5394−5399. and Efficacy. Chem. Rev. 2013, 113, 119−191.
(10) Frank, J. A.; Franquelim, H. G.; Schwille, P.; Trauner, D. Optical (29) König, K. Multiphoton Microscopy in Life Sciences. J. Microsc.
Control of Lipid Rafts with Photoswitchable Ceramides. J. Am. Chem. 2000, 200, 83−104.
Soc. 2016, 138, 12981−12986. (30) Bort, G.; Gallavardin, T.; Ogden, D.; Dalko, P. I. From One-
(11) Frank, J. A.; Moroni, M.; Moshourab, R.; Sumser, M.; Lewin, G. Photon to Two-Photon Probes: “Caged” Compounds, Actuators, and
R.; Trauner, D. Photoswitchable Fatty Acids Enable Optical Control of Photoswitches. Angew. Chem., Int. Ed. 2013, 52, 4526−4537.
TRPV1. Nat. Commun. 2015, 6, 7118. (31) Kantevari, S.; Passlick, S.; Kwon, H.-B.; Richers, M. T.; Sabatini,
(12) Pavlovic, I.; Thakor, D. T.; Vargas, J. R.; Mckinlay, C. J.; Hauke, B. L.; Ellis-Davies, G. C. R. Development of Anionically Decorated
S.; Anstaett, P.; Camuña, R. C.; Bigler, L.; Gasser, G.; Schultz, C.; Caged Neurotransmitters: In Vitro Comparison of 7-Nitroindolinyl-
Wender, P. A.; Jessen, H. J. Cellular Delivery and Photochemical and 2-(p-Phenyl-o-nitrophenyl)propyl-Based Photochemical Probes.
Release of a Caged Inositol-Pyrophosphate Induces PH-Domain ChemBioChem 2016, 17, 953−961.
Translocation in cellulo. Nat. Commun. 2016, 7, 10622. (32) Palma-Cerda, F.; Auger, C.; Crawford, D. J.; Hodgson, A. C. C.;
(13) Broichhagen, J.; Frank, J. A.; Trauner, D. A Roadmap to Success Reynolds, S. J.; Cowell, J. K.; Swift, K. A. D.; Cais, O.; Vyklicky, L.;
in Photopharmacology. Acc. Chem. Res. 2015, 48, 1947−1960. Corrie, J. E. T.; Ogden, D. New Caged Neurotransmitter Analogs
(14) Stein, M.; Middendorp, S. J.; Carta, V.; Pejo, E.; Raines, D. E.; Selective for Glutamate Receptor Sub-Types Based on Methoxyni-
Forman, S. A.; Sigel, E.; Trauner, D. Azo-Propofols: Photochromic troindoline and Nitrophenylethoxycarbonyl Caging Groups. Neuro-
Potentiators of GABAA Receptors. Angew. Chem., Int. Ed. 2012, 51, pharmacology 2012, 63, 624−634.
10500−10504. (33) Xu, C.; Webb, W. W. Measurement of Two-photon Excitation
(15) Hagen, V.; Bendig, J.; Frings, S.; Eckardt, T.; Helm, S.; Reuter, Cross Sections of Molecular Fluorophores with Data from 690 to 1050
D.; Kaupp, U. B. Highly Efficient and Ultrafast Phototriggers for nm. J. Opt. Soc. Am. B 1996, 13, 481−491.
cAMP and cGMP by Using Long-Wavelength UV/Vis-Activation. (34) Albota, M. A.; Xu, C.; Webb, W. W. Two-photon Fluorescence
Angew. Chem., Int. Ed. 2001, 40, 1045−1048. Excitation Cross Sections of Biomolecular Probes from 690 to 960 nm.
(16) Menge, C.; Heckel, A. Coumarin-Caged dG for Improved Appl. Opt. 1998, 37, 7352−7356.
Wavelength-Selective Uncaging of DNA. Org. Lett. 2011, 13, 4620− (35) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid
4623. Density Functionals with Damped Atom-Atom Dispersion Correc-
(17) Rodrigues-Correia, A.; Weyel, X. M. M.; Heckel, A. Four Levels tions. Phys. Chem. Chem. Phys. 2008, 10, 6615−6620.
of Wavelength-Selective Uncaging for Oligonucleotides. Org. Lett. (36) Schaefer, A.; Horn, H.; Ahlrichs, R. Fully Optimized Contracted
2013, 15, 5500−5503. Gaussian-Basis Sets for Atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571−
(18) Sinha, D. K.; Neveu, P.; Gagey, N.; Aujard, I.; Benbrahim- 2577.
Bouzidi, C.; Le Saux, T.; Rampon, C.; Gauron, C.; Goetz, B.; (37) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Dubruille, S.; Baaden, M.; Volovitch, M.; Bensimon, D.; Vriz, S.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
Jullien, L. Photocontrol of Protein Activity in Cultured Cells and B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.
Zebrafish with One- and Two-Photon Illumination. ChemBioChem P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
2010, 11, 653−663. Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,
(19) Herzig, L.-M.; Elamri, I.; Schwalbe, H.; Wachtveitl, J. Light- T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.;
induced Antibiotic Release from a Coumarin-caged Compound on the Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin,
Ultrafast Timescale. Phys. Chem. Chem. Phys. 2017, 19, 14835−14844. K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.;
(20) Ohtsuki, T.; Kanzaki, S.; Nishimura, S.; Kunihiro, Y.; Sisido, M.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.;
Watanabe, K. Phototriggered Protein Syntheses by Using (7- Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.;
Diethylaminocoumarin-4-yl)methoxycarbonyl-caged aminoacyl Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.;
tRNAs. Nat. Commun. 2016, 7, 12501. Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.;
(21) Schade, B.; Hagen, V.; Schmidt, R.; Herbrich, R.; Krause, E.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador,
Eckardt, T.; Bendig, J. Deactivation Behavior and Excited-State P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .;
Properties of (Coumarin-4-yl)methyl Derivatives. 1. Photocleavage Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09,
of (7-Methoxycoumarin-4-yl)methyl-Caged Acids with Fluorescence revision D.01; Gaussian, Inc.: Wallingford, CT, 2013.
Enhancement. J. Org. Chem. 1999, 64, 9109−9117. (38) Buhr, F.; Kohl-Landgraf, J.; tom Dieck, S.; Hanus, C.;
(22) Schmidt, R.; Geissler, D.; Hagen, V.; Bendig, J. Kinetics Study of Chatterjee, D.; Hegelein, A.; Schuman, E. M.; Wachtveitl, J.;
the Photocleavage of (Coumarin-4-yl)methyl Esters. J. Phys. Chem. A Schwalbe, H. Design of Photocaged Puromycin for Nascent
2005, 109, 5000−5004. Polypeptide Release and Spatiotemporal Monitoring of Translation.
(23) Schmidt, R.; Geissler, D.; Hagen, V.; Bendig, J. Mechanism of Angew. Chem., Int. Ed. 2015, 54, 3717−3721.
(39) Slavov, C.; Bellakbil, N.; Wahl, J.; Mayer, K.; Rück-Braun, K.;
Photocleavage of (Coumarin-4-yl)methyl Esters. J. Phys. Chem. A
Burghardt, I.; Wachtveitl, J.; Braun, M. Ultrafast Coherent Oscillations
2007, 111, 5768−5774.
Reveal a Reactive Mode in the Ring-opening Reaction of Fulgides.
(24) Gatterdam, V.; Ramadass, R.; Stoess, T.; Wachtveitl, J.; Heckel,
Phys. Chem. Chem. Phys. 2015, 17, 14045−14053.
A.; Tampé, R. Three-dimensional Protein Networks Guided by Two-
(40) Slavov, C.; Hartmann, H.; Wachtveitl, J. Implementation and
Photon Activation. Angew. Chem., Int. Ed. 2014, 53, 5680−5684.
Evaluation of Data Analysis Strategies for Time-Resolved Optical
(25) Brown, E. B.; Shear, J. B.; Adams, S. R.; Tsien, R. Y.; Webb, W.
Spectroscopy. Anal. Chem. 2015, 87, 2328−2336.
W. Photolysis of Caged Calcium in Femtoliter Volumes Using Two-
Photon Excitation. Biophys. J. 1999, 76, 489−499.
(26) Ellis-Davies, G. C. R. Caged Compounds: Photorelease
Technology for Control of Cellular Chemistry and Physiology. Nat.
Nat. Methods 2007, 4, 619−628.
(27) Fichte, M. A. H.; Weyel, X. M. M.; Junek, S.; Schäfer, F.;
Herbivo, C.; Goeldner, M.; Specht, A.; Wachtveitl, J.; Heckel, A.
Caged Biomolecules Three-Dimensional Control of DNA Hybrid-
ization by Orthogonal Two-Color Two-Photon Uncaging. Angew.
Chem., Int. Ed. 2016, 55, 8948−8952.
(28) Klán, P.; Šolomek, T.; Bochet, C. G.; Blanc, A.; Givens, R.;
Rubina, M.; Popik, V.; Kostikov, A.; Wirz, J. Photoremovable

1453 DOI: 10.1021/acs.jpclett.7b03364


J. Phys. Chem. Lett. 2018, 9, 1448−1453

You might also like