Comparison of Complexation-Induced pKa Shifts in the Ground and

Nov 23, 2017 - Comparison of Complexation-Induced pKa Shifts in the Ground and Excited States of Dyes as Well as Different Macrocyclic Hosts and Their...
0 downloads 7 Views 919KB Size
Subscriber access provided by READING UNIV

Article a

Comparison of Complexation-Induced pK Shifts in the Ground and Excited State of Dyes as Well as Different Macrocyclic Hosts and Their Manifestation in Host-Retarded Excited-Dye Deprotonation Alexandra Irina Lazar, Jana Rohacova, and Werner M. Nau J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.7b10651 • Publication Date (Web): 23 Nov 2017 Downloaded from http://pubs.acs.org on November 24, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry B is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Comparison of Complexation-Induced pKa Shifts in the Ground and Excited State of Dyes as Well as Different Macrocyclic Hosts and Their Manifestation in Host-Retarded Excited-Dye Deprotonation Alexandra I. Lazar, Jana Rohacova,† Werner M. Nau* Department of Life Sciences and Chemistry, Jacobs University Bremen, Campus Ring 1, 28759 Bremen, Germany

ABSTRACT: 1-Aminopyrene and 1-naphthylamine-5-sulfonic acid were converted to the putrescine (1,4-diaminobutane)-substituted derivatives (dyes 1 and 2). The diaminobutyl anchor serves as a common binding motive for cation-receptor macrocycles such as cucurbit[n]urils (n = 6-8) and p-sulfonatocalix[4]arene. When protonated, they are prone to undergo a rapid deprotonation in their excited state to result in fluorescence from the unprotonated form (Förster cycle). The deprotonation can be suppressed by complexation with cation-receptor macrocycles, which allows the fluorescence of the locally excited (protonated) state to be dramatically

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 37

enhanced (factor 12 for dye 1 and 83 for dye 2). This host-retarded excited-dye deprotonation is a direct consequence of the previously established complexation-induced pKa shifts which dyes undergo upon binding to a macrocyclic host. The data set allows also a systematic comparison of complexation-induced pKa shifts in the ground and excited state of a dye. The trends are comparable, which suggests that structural factors, that is, the geometry of the host-guest complexes, determine the magnitude of the shifts. In regard to the magnitude of the absolute pKa shifts on the size of the macrocycles, we observe for dye 2 that the complexation-induced pKa shifts decrease as the portal becomes larger for the CBn series.

INTRODUCTION The study of the interactions of fluorescent dyes with water-soluble macrocyclic host molecules has developed into an important area of supramolecular chemistry.1-32 Upon complexation, the fluorescent dye frequently changes significantly, rarely dramatically, its fluorescence properties, such that both its complexation as well as its decomplexation can be followed with high sensitivity. This fluorescence response provides a robust method to i) measure and compare the binding propensities of different macrocycles at low dye concentrations,33-35 ii) explore the physical properties of the inner cavities of macrocycles by means of solvatochromic effects,10, 3640

iii) develop refined fluorescent switches for advanced supramolecular architectures6, 10, 13, 17, 41-

43

or molecular logic gates,23, 44-47 and iv) monitor the reversible complexation of analytes to the

macrocycles through indicator displacement strategies.48-54 The latter method, because it can be both miniaturized and scaled up to high-throughput screening format, has enormous application potential in the environmental, biological, and food sciences. It also forms the conceptual basis for selective fluorescent sensing ensembles operational in aqueous solution.17,

35, 54-62

A large

ACS Paragon Plus Environment

2

Page 3 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

fluorescence response of the dye upon complexation is invariably desirable. Hence, the understanding of the mechanistic details, which lead to the fluorescence changes, and the quest for new mechanisms present a timeless endeavor. We now designed new dyes (1 and 2, Scheme 1) of the “anchor dye” type,6 which combine a common recognition motif for cation-receptor hosts with a basic nitrogen as protonation site, such that ground-state pKa shifts can be systematically analyzed for a series of hosts, in particular cucurbit[n]urils (CBn, n = 6-8)63-70 and p-sulfonatocalix[4]arene (CX4).57 In addition, we selected chromophores that are prone to undergo a Förster cycle, which, through the concomitant observation of unprotonated as well as protonated dye fluorescence, allow also the determination of excited-state pKa shifts. To our initial surprise, the excited-state deprotonation (Förster cycle) of dyes 1 and 2 was strongly suppressed by complexation. Closer inspection of the combination of the two phenomena (pKa shifts and Förster cycle) showed that the slower deprotonation kinetics, which manifests itself through an unexpectedly high fluorescence of the protonated complexed dye, is inevitably linked to the thermodynamics of the pKa shifts. The peculiar photophysical behavior can be considered as a distinct mechanism for switching on and off the fluorescence of dyes, which can be utilized for sensing. Effects of CB7 on excited-state protonation phenomena of dyes have been noted in different contexts and controversially discussed.10,

71-74

Macartney and coworkers investigated 2-

aminoanthracene with CB771 and interpreted the phenomenon as a thermodynamic excited-state effect, that is, a much larger excited-state shift (∆pKa* ca. 10 units) than the ground state one (∆pKa ca. 3 units).71 Scrutiny of their analysis revealed, however, an inconsistent treatment (see Supporting Information) and recalculation affords the same value for both, ground- and excitedstate pKa. The analysis of Macartney formed the basis of Thomas and Bohne’s work.28

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 37

Hazra and coworkers also claimed a much larger ∆pKa* value of 8.1 units than the ∆pKa value of 0.6 units for topotecan with CB7.72 Inspection of their experimental data showed that they assigned the inflection point of the steady-state fluorescence titrations to the pKa* values; such direct assignment is not justified, because fluorescence titration curves also report on pKa values.21, 75 Basílio et al. inspected the effect of CB7 on two dyes, namely 4-methyl-7-hydroxyflavylium and β-naphthol, and found the apparent ∆pKa'* values to be 1.2 in the case of 4-methyl-7hydroxyflavylium and undetermined for β-naphthol, because the excited-state proton transfer (ESPT) was completely suppressed by CB7 in the latter case.73 The authors did not provide a comprehensive interpretation but they concluded instead that: “This can be due to effective suppression of the ESPT or to a shift of pKa* to higher values”.73 Our present study for dyes 1 and 2 – as well as one of our previous studies10 and the recalculation of the measurements by Macartney (see Supporting Information) – show that the pKa shifts are actually comparable in the ground and excited-state. Moreover, the observed apparent stabilization of the protonated complexed dye is a consequence of the reduction in deprotonation rate. This slow kinetics introduces a “metastability” of the excited-state, allowing it to become fluorescent and observable even at pH conditions in which the unprotonated form would otherwise be thermodynamically favored.

ACS Paragon Plus Environment

4

Page 5 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Scheme 1. Chemical structures of dyes 1 and 2 (in their protonated DH+ form) as well as the employed macrocycles. First deprotonation sites of the two dyes (to afford the D form) are indicated by arrows.

RESULTS Synthesis. Compound 1 was synthesized by coupling (reductive amination)76 of 1-aminopyrene with 4-phthalimido-butyraldehyde77 and final deprotection of the 4-aminobutyl group with an overall yield of 46% (see Supporting Information). Compound 2 was synthesized analogously in similar overall yield (50%). The diaminobutyl chain of dyes 1 and 2 infers a high affinity for cation-receptor macrocycles such as CBn and CX4, which we have previously used as model

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 37

hosts to study complexation-induced pKa shifts (host-induced guest protonation) with other dyes.3, 6-7, 10, 19, 78-79

Table 1. Thermodynamic dataa of the protonated and unprotonated dye 1 Host

KDH (M–1)

KD (M–1)

pKa'

∆pKa

2.9b [2.8b,c]

None

pKa'*

∆pKa'*

–5.8d

CB6

2.0 × 108

2.2 × 105

5.7

2.8

–3.6

2.2

CB7

1.3 × 104

150

5.5

2.6

–3.2

2.6

CB8

6.3 × 106

4.8 × 104

6.6

3.7

–2.1

3.7

CX4

2.8 × 104

200

6.3

3.4

–2.4

3.4

a

Error in pKa values is 0.1. from ref. 80. d pKa*.

b

pKa. c Literature value for parent chromophore (1-aminopyrene)

Table 2. Thermodynamic dataa of the protonated and unprotonated dye 2 Host

KDH (M–1)

KD (M–1)

pKa'

∆pKa

2.9b [3.7b,c]

None

pKa'*

∆pKa'*

–13.5d

CB6

1.3 × 108

6.4 × 104

6.2

3.3

–10.9

2.7

CB7

3.1 × 104

230

5.1

2.2

–11.5

2.0

CB8

2.2 × 104

2.0 × 103

4.0

1.1

–12.9

0.7

CX4

1.8 × 103

1.1 × 103

3.2

0.3

–12.9

0.7

a

Error in pKa values is 0.1. b pKa. c Literature value for the parent chromophore (1naphthylamine-5-sulfonic acid) from ref. 81. d pKa*.

ACS Paragon Plus Environment

6

Page 7 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. Absorption (black) and emission (red) spectra (λexc = 346 nm for 1 and λexc = 301 nm for 2, the respective isosbestic points) of dye 1 (a) and 2 (b) in their protonated (pH 1, solid line) and unprotonated (pH 7, dashed line) forms.

Physicochemical and photophysical properties. In contrast to the terminal alkylammonium groups of dyes 1 and 2 (pKa ca. 10), their arylamino groups are unprotonated at neutral pH, but can be protonated at strongly acidic pH, as can be nicely followed through the interconversion of the broad long-wavelength absorption band at 360 and 335 nm, respectively, characteristic for

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 37

the intramolecular charge-transfer transition (ICT, D→D*) of the amino form of the dyes, to a structured short-wavelength absorption at 340 and 283 nm, respectively, arising from the transition of the ammonium dyes to their locally excited states (LE, DH+→DH+*). The corresponding UV and fluorescence titrations (see Supporting Information) provide the same pKa values of the arylammonium groups of 1 and 2 (2.9, see Tables 1 and 2), falling slightly but systematically below the pKa values of the parent primary amino chromophores (3.4 and 3.7),8182

an effect which is due to the electrostatic influence of the terminal ammonium group.

However, regardless which state is being populated (D* or DH+*), and regardless at which pH the excitation is performed (pH 0-7), the fluorescence remains effectively the same for both dyes: a broad ICT-type emission from the D* state centered at 454 nm for 1 and at 507 nm for 2, Figure 1. This demonstrates that even when the DH+* state is populated at very low pH, it undergoes a rapid deprotonation due to a dramatically lowered excited-state pKa* value (DH+*→D*→D + hʋ). Such a phenomenon is well known as the Förster cycle and resembles the photophysics of the parent 1-aminopyrene (pKa* ca. –5 in aqueous media).80,

83-85

Close

inspection of the spectra reveals, in fact, that there is a residual emission from DH+*, but it is very weak (see Figure 1 and Scheme 2a) due to the short lifetime of the LE states (5.20 and 1.50 ns, see Table 3). The concomitant observation of the absorption and fluorescence spectra of both, unprotonated and protonated forms of the dyes, allowed us to determine their excited-state pKa* values (–5.8 and –13.5 for dye 1 and 2, Tables 1 and 2) according to the common Förster cycle, see Supporting Information, which are entropy-uncorrected values. The values resulting from an empirical entropy correction developed by Shizuka86 differ systematically (–4.35 and –5.40 for dye 1 and 2, Supporting Information).

ACS Paragon Plus Environment

8

Page 9 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Scheme 2. Förster cycle for the free (a) and complexed (b) dyes 1 and 2 (D). The bold arrows indicate the faster of two kinetically competitive processes.

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 37

Table 3. Lifetime values (τF) for dyes 1 and 2 as well as their complexes with macrocyclic hosts τF (ns) Host

a c

1H+a

1b

2H+c

2d

None

5.20

4.60

1.50

9.12

CB6

156

4.44

14.3

8.97

CB7

154

4.93

9.47e

10.6e

CB8

135

5.53

8.37

7.45

CX4

16.5

5.35

-

-

λexc = 340 nm, λem = 375 nm, solutions de-aerated with Ar. b λexc = 373 nm, λem = 455 nm. λexc = 280 nm, λem = 330 nm. d λexc = 335 nm, λem = 505 nm. e Average lifetime values

(biexponential decay).

Complexation by cation-receptor macrocycles. The addition of the macrocyclic hosts CB6, CB7, CB8, and CX4 to solutions of dyes 1 and 2 at different pH values caused characteristic changes in their absorption and fluorescence spectra, attributable to the formation of inclusion complexes (Figure 2 and Supporting Information).6 The spectral changes were used to determine the binding constants of the protonated and unprotonated forms of the dyes and their pKa values in the free and complexed forms (Tables 1 and 2, Supporting Information) by means of optical titrations. The corresponding analytical procedures to determine the associated complexationinduced pKa shifts in the ground state have already been abundantly documented in numerous previous studies.2-3, 7, 78, 87-89

ACS Paragon Plus Environment

10

Page 11 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2. (a) Fluorescence emission spectra of 0.5 µM 1 in the absence (dashed line) and presence (solid line) of 0.9 µM CB6, at pH 1.5, λexc = 347 nm; (b) fluorescence emission spectra of 10 µM 2 in the absence (dashed line) and presence (solid line) of 1.6 mM CB7, at pH 1.5,

λexc = 301 nm. Dyes 1 and 2 display a special behavior when their characteristic fluorescence changes were investigated at pH 1.5 (see Figure 2 and Supporting Information). This pH is sufficiently acidic to assure virtually quantitative protonation of both, the complexed and uncomplexed forms of dye 1 and 2, such that variations in the optical spectra cannot be attributed to complexation-

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 37

induced pKa shifts on the ground-state equilibria. Nevertheless, addition of the different macrocyclic hosts led to dramatic changes in the fluorescence spectra, namely an increase in the emission band of the protonated LE state (up to a factor of 83 for the CB6•2 complex) at the expense of a concomitant decrease in emission from the unprotonated ICT state. In view of the Förster cycle, it appeared that the protonated excited state is stabilized by the macrocycle. To evaluate the thermodynamic effect of encapsulation on the stability of the excited states we determined also the pKa'* values according to the Förster cycle (−2 to −13, entropy-uncorrected values, Tables 1 and 2). The entropy-corrected values remained negative and followed the same trend (Supporting Information). Accordingly, the deprotonated states of both dyes remain strongly thermodynamically favored even at the selected acidic pH (1.5). The observed fluorescence response of these two dyes upon macrocyclic complexation must therefore be due to a kinetic inhibition of the deprotonation of the complexed dyes in their excited states. To quantify the kinetic effect of encapsulation on deprotonation, time-resolved fluorescence measurements were performed.28, 73, 90 Indeed, the protonated dyes 1 and 2 in their complexes with CBs, and to a lesser degree in their complexes with CX4, showed drastic enhancements of their LE-state fluorescence lifetimes (Table 3); this reflects a higher kinetic persistence which accounts also for the enhanced fluorescence intensities (Figure 2) and fluorescence quantum yields (φF values in Table 4 for dye 1) corresponding to emission from the protonated LE state. For example, the fluorescence lifetime increased from 5.20 ns for the protonated form of the free dye 1 to 156 ns for its CB6 complex (solutions were de-aerated to eliminate oxygen quenching). Since the radiative decay rates (kr values in Table 4), calculated as the product of the quantum yield and the reciprocal fluorescence lifetime, were little affected, the increase in lifetime can be attributed to a reduction of the non-radiative decay rate upon complexation, namely a

ACS Paragon Plus Environment

12

Page 13 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

suppression of the deprotonation of the protonated excited-state. For comparison, the fluorescence quantum yield of the unprotonated ICT state, determined at higher pH values, was insignificantly affected by the macrocyclic complexation (Table 4). Table 4. Quantum yields (φF), excited-state deprotonation rates (kdeprot), and radiative fluorescence decay rates (kr) of dye 1a with macrocyclic hosts DH+b (LE state) Host

φF None 0.02

Dc (ICT state)

kdeprot d (106 s-1) kr e (106 s–1) 190 (2200)f

φF

kr e (108 s–1)

3.9

0.67

1.46

CB6g 0.55

2.9

3.5

0.73

1.66

CB7h 0.33

4.4

2.1

0.60

1.22

CB8i 0.43

4.2

3.2

0.61

1.11

2.4

0.67

1.26

CX4j 0.04

58

a

[1] = 5 µM. b pH = 1.5. c pH = 7.5. d Approximate values calculated as described in the Supporting Information. e Calculated with kr = φF/τF, with τF values from Table 3. f The kdeprot value of 1-aminopyrene, the parent chromophore of dye 1, from ref. 80. g [CB6] = 20 µM. h [CB7] = 100 µM. i [CB8] = 20 µM. j [CX4] = 200 mM.

The ability to follow the deprotonation event (of the excited-state) in real time offers the opportunity to substantiate the thermodynamic effect that macrocycles display on the pKa values of dyes in kinetic terms, i.e., on the rate constants of (de)protonation (Scheme 2b). In detail, the fluorescence lifetimes (Table 3), along with the independently determined fluorescence quantum yields from the LE-state emission (Table 4), allow us to quantify the excited-state deprotonation rates (kdeprot) with the assumption that the non-radiative decay of the protonated form occurs exclusively by deprotonation. This analysis, which is detailed in the Supporting Information, affords kdeprot values which do indeed decrease by up to two orders of magnitude upon

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 37

complexation (Table 4). Consideration of additional non-radiative decay pathways would affect the absolute kdeprot values, but is unlikely to change the pronounced relative trends. By further assuming a constant protonation rate (Scheme 2), an independent estimate for the pKa* values can be obtained (“kinetic” calculation of the pKa* values, Table S3, Supporting Information). Although these were systematically higher than the pKa* values derived from the Förster cycle and the empirical entropy correction (Table S3, Supporting Information), the resulting values remain below the selected pH value, confirming that the protonated excited state of dyes 1 and 2 are not energetically more stable than the unprotonated excited forms but kinetically more persistent due to a retarded deprotonation in the complex. The combined optical spectroscopic data lead us to propose the state diagram in Scheme 2 as a summarized result, in which eight states (ground and excited, complexed and uncomplexed, protonated and unprotonated) are interrelated. The corresponding free-energy diagram is shown in Scheme 3, which reveals that the origin of the observed fluorescence enhancement of the protonated form accompanying the complexation is not related to the fact that the protonated form becomes more stable than the unprotonated form, but rather that its interconversion to the unprotonated form is kinetically hindered. The protonated state becomes “metastable” on the excited-state time scale.

ACS Paragon Plus Environment

14

Page 15 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Scheme 3. Free-energy diagram for the protonation and complexation equilibria of dye 2, depicted for CB6 as host, the experimentally employed pH value of 1.5, and a typical host concentration of 20 µM.

Exploratory chemosensing applications. While the present study is devoted to photophysical aspects, it should be added that the fluorescence response of dyes 1 and 2 to macrocyclic complexation can be readily implemented into chemosensing applications. To illustrate, we have successfully employed the CB6•1 reporter pair for sensing of cadaverine as a reaction product of the enzyme lysine decarboxylase, see Supporting Information (Figure S32); cadaverine binds tightly to CB6 and its presence or formation reverts the fluorescence response associated with the inclusion of 1 into the CB6 cavity. Independently, the CB6•2 reporter pair has already been communicated to serve as gas sensor,60 since simple alkanes have a high affinity for this host and compete with 2 for inclusion into the CB6 cavity.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 37

DISCUSSION This work provides a contribution to the understanding of the interaction of fluorescent dyes with supramolecular assemblies and applications of the resulting macrocycle•dye complexes. The novelty of our study with dyes 1 and 2 is twofold. It comprises (i) a systematic comparison of ground- and excited-state pKa shifts and their variation for different hosts and (ii) an investigation of the effects of macrocyclic complexation on the deprotonation kinetics. Compounds 1 and 2 are designer dyes in which classical fluorophores, namely 1-aminopyrene and 5-aminonaphthalene-sulfonate, were decorated with an aminobutyl tether. The latter serves as a common anchor to facilitate the complexation with a series of macrocyclic cation-receptor hosts: three homologous cucurbiturils (CB6, CB7, CB8) and p-sulfonato-calix[4]arene (CX4). Drawing from a previous exploratory utilization of this anchor strategy,6 the associated systematic investigation allowed us to obtain the first structure-activity relationship of groundstate pKa shifts for different macrocyclic complexes of the same dye, complementing the scattered data on pKa shifts available for different host•dye combinations in the literature. The state of affairs of pKa shifts induced by macrocyclic complexation has been first summarized in 20062, 87 and several more examples have become available in the following decade.6-7, 10, 19, 21-23, 28, 74, 78, 91-94

As can be seen from Table 1 and Table 2, all four receptors bind to the protonated forms of dyes 1 and 2 with high affinities (KDH ≈ 103 – 108 M-1) which confirms the validity of our anchor design. The selected putrescine is known to have the highest affinity for CB6 among the cucurbituril series, which is also reflected by the experimental KDH obtained for 1 and 2. The corresponding pH titrations, followed by both UV-vis and fluorescence spectroscopy, afforded the pKa values of the complexed dyes, which were markedly higher than those of the free dyes,

ACS Paragon Plus Environment

16

Page 17 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

as expected. The hosts “assist” the protonation of the dyes because of their known preference to bind cations.3, 6-7, 10, 19, 23, 26, 32, 34, 43, 46, 87, 94-96 The pKa shifts induced by the macrocycles (∆pKa) reached up to +4 pKa units, which falls in the reported range.87 Moreover, the rationally selected chromophores undergo a Förster cycle, that is, they have a known propensity to deprotonate from their excited state as a consequence of a lower pKa* compared to their pKa. The dominant fluorescence stems from the unprotonated excited state, that is, it occurs with a large Stokes shift at longer wavelengths. The pKa* values can be simply determined from the spectral properties (0–0 transitions in absorption and emission spectra) for the free and complexed forms (Tables 1 and 2). Compared to the pKa values, the pKa* values of dyes 1 and 2 become strongly negative (decreased by ca. 9 units for dye 1 and 16 units for dye 2), reflecting the high acidity of their excited states. The pKa* values of the macrocycle•dye complexes are also 9 – 16 units lower than their pKa counterparts, but the absolute values of both, pKa* and pKa, are offset to positive values by up to +4 units due to the macrocyclic complexation. In fact, through the Förster cycle it becomes possible to extract the complexationinduced pKa shifts in the excited state and compare them with these in the ground state (compare ∆pKa* with ∆pKa in Tables 1 and 2). The experimental data for dyes 1 and 2 allow us to derive the first structure-activity relationships on pKa shifts induced by macrocyclic complexation: (I) Complexation-induced pKa shifts are comparable in the ground and excited state. This is a strong indication that the pKa shifts are due to “simple” geometrical (steric) reasons, because no large co-conformational changes are expected during electronic excitation and/or the short singlet excited-state lifetimes. Practically, the pKa shifts in the ground state are more important, but knowledge on the excited-

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 37

state behavior is theoretically relevant because fluorescence is a popular method for pKa shift determination. In regard to pKa shifts in dependence with the type of macrocycle, dye 1 proved to be a less fortunate model because its hydrophobic pyrene moiety competed with the diaminobutyl anchor for binding to macrocycles with sufficiently large or flexible cavities, such as CB8 and CX4. Evidence for this property was obtained by 1H NMR complexation-induced chemical shifts and solvatochromic shifts (see Supporting Information). This resulted in different co-conformations and positions of the protonation site such that trends in dependence on the type of macrocycle were inconclusive for dye 1. This complication does not apply for dye 2 which carries a negatively charged sulfonato group at the naphthalene residue which suppresses binding to the selected cation-receptor macrocycles, confirmed as well by 1H NMR. Consequently, we used the thermodynamic data for dye 2 to advance a second structure-activity relationship: (II) Within the cucurbituril series (n = 6-8), when the inclusion mode is comparable, complexation-induced pKa shifts decrease as the portal becomes larger. This observation is consistent with expectations based on Coulombic attraction (ECoulomb ∝ n/RCBn), which increases with the number of interaction sites (n), but decreases with the distance between the charges (RCBn), i.e., the protonation site and the carbonyl portal oxygens. A simple calculation with the assumption of a centrosymmetric charge in vacuum and the pKa shift of CB6 arbitrarily set to 3.0 (similar to the experimental values, see Table 2), predicts relative pKa shifts of 2.5 for CB7 and 2.3 for CB8, and, thereby, the correct trend. The fact that the experimental decrease is more pronounced than predicted can be attributed to the at least two reasons, the first being a proximate interaction of the protonation site with 1-2 carbonyl groups and more distal ones with the others (rather than a centrosymmetric geometry), the second one being an increased charge screening due to the better

ACS Paragon Plus Environment

18

Page 19 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

hydration enabled by the larger portals; recall that the relative dielectric constant, which is large for water, enters as a factor in the denominator of the Coulombic energy term. Even though our study allows us to understand the dependence of the complexation-induced pKa shifts on the electronic state and the macrocycle size, unknowns remain. These extend to the large variations of experimentally observed pKa shifts on the precise structure of the dyes, which have been noted before.19,

87

The large differences of the complexation-induced pKa shifts

between dye 1 and 2 with the larger macrocycles (CB8 and CX4) can be attributed to the different dye immersion pattern (see above). However, even when a common anchor motif is there to assure that the protonation site is always an arylamino nitrogen such as in the CB6 complexes of dyes 1 and 2 as well as a previously studied carbazole dye6 – the corresponding pKa shifts differ significantly, from 2.8 to 4.5 (Tables 1 and 2 and ref.6). This suggests that minute variations in the precise positioning of the protonated site inside macrocycles − unavoidable when attaching different chromophores – have a strong impact on the absolute pKa shifts achieved. While nature has optimized this subtle interplay and can realize pKa shifts up to 5 units in proteins,87, 97-101 the rational design of synthetic host-guest complexes exceeding this benchmark remains a challenge. Beside the derived structure-activity relationships, the project afforded an element of surprise and practical importance; this refers phenomenologically to the unexpected presence of the protonated-state emission in the complexes and kinetically to a dramatically increased excitedstate lifetime of the protonated state. This means that the macrocycles have a peculiar “photophysical function” – they apparently suppress the Förster cycle and induce a kinetic metastability of the protonated dye, which is otherwise known to undergo an efficient deprotonation to the excited unprotonated form, from where fluorescence conventionally predominates.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 37

The suppression of the Förster cycle, affected by the addition of cation receptors, in general, and cucurbiturils, in particular, establishes a distinct mechanism through which the fluorescent properties of dyes can be modulated. Such a selective chemoresponse is highly desirable for sensing applications and is superior to a less specific quenching or microenvironmental response of fluorescent dyes, such as the more commonly encountered dependencies on polarity, viscosity or confinement.95 It also complements the use of ground-state complexation-induced pKa shifts of dyes as a previously established thermodynamical tool for optical response; 3, 6-7, 10, 19, 23, 26, 32, 34, 43, 46, 74, 87, 92, 94-96

these are operationally limited to pH values in between the pKa values of the

complexed and uncomplexed dye. In contrast, the kinetic phenomenon of host-retarded excitedstate deprotonation is also observed outside this pH range (Schemes 2 and 3).

CONCLUSIONS The investigation of the anchor dyes 1 and 2 with a series of macrocyclic host molecules has exposed the first two structure-activity relationships on complexation-induced pKa shifts and thereby contributes to the partial understanding of the more than 100 experimental data that have been reported over the past 15 years.3, 6-7, 13-14, 18-19, 23, 26, 29, 50, 59, 62, 70-71, 74, 87, 91-92, 94-95, 102-107 In addition, a new mechanism, host-retarded excited-state deprotonation, which complements the abundantly documented host-assisted (ground-state) dye protonation3, 6, 18-19, 21, 23, 32, 34, 43, 46, 74, 87, 92, 94-96

as a tool to modulate the fluorescence, has been scrutinized. It rests on the stabilization of

the excited dye whose otherwise efficient deprotonation pathway is kinetically hindered in its complexed form. The new mechanism and the design approach can be implemented into a broad range of applications, such as indicator displacement assays.53, 108 In fact, both dyes have already been used in enzyme assays (dye 1, see Results) and for chemosensing of gases60 (dye 2). From a

ACS Paragon Plus Environment

20

Page 21 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

photophysical point of view, the complexation of such dyes by macrocycles provides a unique approach to dramatically “shift” Förster cycles without changing the solvent and this opportunity can be exploited, among others, to access otherwise elusive excited states, for example, for spectroscopic characterization.

ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI: Materials and methods, pKa* calculations, synthesis, application example (PDF).

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]; Tel.: +49 421 200 3233 ORCID Alexandra I. Lazar: 0000-0003-1854-7920 Jana Rohacova: 0000-0003-4828-6231 Werner M. Nau: 0000-0002-7654-6232 Present Addresses †

(J.R.) Teva Czech Industries S.R.O., Ostravská 29, 747 70 Opava, Czech Republic.

Author Contributions A. I. L. and J. R. contributed equally.

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 37

Notes The authors declare no competing financial interests.

ACS Paragon Plus Environment

22

Page 23 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ACKNOWLEDGMENT We would like to acknowledge the financial support of the Deutsche Forschungsgemeinschaft (DFG, NA-686/11-1).

REFERENCES (1)

Nau, W. M.; Mohanty, J. Taming Fluorescent Dyes with Cucurbituril. Int. J. Photoenergy

2005, 7, 133-141. (2)

Mohanty, J.; Bhasikuttan, A. C.; Nau, W. M.; Pal, H. Host−Guest Complexation of

Neutral Red with Macrocyclic Host Molecules: Contrasting pKa Shifts and Binding Affinities for Cucurbit[7]uril and β-Cyclodextrin. J. Phys. Chem. B 2006, 110, 5132-5138. (3)

Koner, A. L.; Nau, W. M. Cucurbituril Encapsulation of Fluorescent Dyes. Supramol.

Chem. 2007, 19, 55-66. (4)

Wagner, B. D. Recent Applications of Host-Guest Inclusion in Fluorescence-Based Trace

Analysis. Curr. Anal. Chem. 2007, 3, 183-195. (5)

Nau, W. M.; Hennig, A.; Koner, A. L. Squeezing Fluorescent Dyes into Nanoscale

Containers—The Supramolecular Approach to Radiative Decay Engineering; In Fluorescence of Supermolecules, Polymers, and Nanosystems; Berberan-Santos, M. N., Ed. Springer: Berlin, Heidelberg, Germany, 2008; Vol. 4, pp. 185-211. (6)

Praetorius, A.; Bailey, D. M.; Schwarzlose, T.; Nau, W. M. Design of a Fluorescent Dye

for Indicator Displacement from Cucurbiturils: A Macrocycle-Responsive Fluorescent Switch Operating through a pKa Shift. Org. Lett. 2008, 10, 4089-4092.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(7)

Page 24 of 37

Saleh, N. i.; Koner, A. L.; Nau, W. M. Activation and Stabilization of Drugs by

Supramolecular pKa Shifts: Drug-Delivery Applications Tailored for Cucurbituri. Angew. Chem., Int. Ed. 2008, 47, 5398-5401. (8)

Shaikh, M.; Mohanty, J.; Bhasikuttan, A. C.; Uzunova, V. D.; Nau, W. M.; Pal, H. Salt-

Induced Guest Relocation from a Macrocyclic Cavity into a Biomolecular Pocket: Interplay between Cucurbit[7]uril and Albumin. Chem. Commun. 2008, 3681-3683. (9)

Montes-Navajas, P.; Garcia, H. Complexes of Basic Tricyclic Dyes in Their Acid and

Basic Forms with Cucurbit[7]uril: Determination of pKa and Association Constants in the Ground and Singlet Excited State. J. Photochem. Photobiol., A 2009, 204, 97-101. (10) Shaikh, M.; Dutta Choudhury, S.; Mohanty, J.; Bhasikuttan, A. C.; Nau, W. M.; Pal, H. Modulation of Excited-State Proton Transfer of 2-(2’-Hydroxyphenyl)benzimidazole in a Macrocyclic Cucurbit[7]uril Host Cavity: Dual Emission Behavior and pKa Shift. Chem. - Eur. J. 2009, 15, 12362-12370. (11) Wagner, B. D. The Use of Coumarins as Environmentally-Sensitive Fluorescent Probes of Heterogeneous Inclusion Systems. Molecules 2009, 14, 210-237. (12) Zhang, H.; Wang, Q.; Liu, M.; Ma, X.; Tian, H. Switchable V-Type [2]Pseudorotaxanes. Org. Lett. 2009, 11, 3234-3237. (13) Sinha, M. K.; Reany, O.; Parvari, G.; Karmakar, A.; Keinan, E. Switchable Cucurbituril− Bipyridine Beacons. Chem. - Eur. J. 2010, 16, 9056-9067. (14) Wyman, I. W.; Macartney, D. H. Host−Guest Complexations of Local Anaesthetics by Cucurbit[7]uril in Aqueous Solution. Org. Biomol. Chem. 2010, 8, 247-252.

ACS Paragon Plus Environment

24

Page 25 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(15) Biczók, L.; Wintgens, V.; Miskolczy, Z.; Megyesi, M. Fluorescence Response of Alkaloids and DAPI on Inclusion in Cucurbit[7]uril: Utilization for the Study of the Encapsulation of Ionic Liquid Cations. Isr. J. Chem. 2011, 51, 625-633. (16) Dsouza, R. N.; Pischel, U.; Nau, W. M. Complexation of Fluorescent Dyes by Macrocyclic Hosts; In Supramolecular Effects on Photochemical and Photophysical Processes; V. Ramamurthy, Y. I., Ed. John Wiley & Sons: Hoboken, NJ, USA, 2011. (17) Parvari, G.; Reany, O.; Keinan, E. Applicable Properties of Cucurbiturils. Isr. J. Chem. 2011, 51, 646-663. (18) Saleh, N. i.; Meetani, M. A.; Al-Kaabi, L.; Ghosh, I.; Nau, W. M. Effect of Cucurbit[n]urils on Tropicamide and Potential Application in Ocular Drug Delivery. Supramol. Chem. 2011, 23, 650-656. (19) Ghosh, I.; Nau, W. M. The Strategic Use of Supramolecular pKa Shifts to Enhance the Bioavailability of Drugs. Adv. Drug Delivery Rev. 2012, 64, 764-783. (20) Muddana, H. S.; Varnado, C. D.; Bielawski, C. W.; Urbach, A. R.; Isaacs, L.; Geballe, M. T.; Gilson, M. K. Blind Prediction of Host−Γuest Binding Affinities: A New SAMPL3 Challenge. J. Comput. Aided Mol. Des. 2012, 26, 475-487. (21) Gavvala, K.; Sengupta, A.; Hazra, P. Modulation of Photophysics and pKa Shift of the Anti-cancer Drug Camptothecin in the Nanocavities of Supramolecular Hosts. ChemPhysChem 2013, 14, 532-542. (22) Barooah, N.; Sundararajan, M.; Mohanty, J.; Bhasikuttan, A. C. Synergistic Effect of Intramolecular Charge Transfer toward Supramolecular pKa Shift in Cucurbit[7]uril Encapsulated Coumarin Dyes. J. Phys. Chem. B 2014, 118, 7136-7146.

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 37

(23) Barooah, N.; Mohanty, J.; Pal, H.; Bhasikuttan, A. C. Cucurbituril-Induced Supramolecular pKa Shift in Fluorescent Dyes and Its Prospective Applications. Proc. Natl. Acad. Sci., India, Sect. A 2014, 84, 1-17. (24) De Filippo, C. C.; Tang, H.; Ravotto, L.; Bergamini, G.; Salice, P.; Mba, M.; Ceroni, P.; Galoppini, E.; Maggini, M. Synthesis and Electronic Properties of 1,2-Hemisquarimines and Their Encapsulation in a Cucurbit[7]uril Host. Chem. - Eur. J. 2014, 20, 6412-6420. (25) Gavvala, K.; Koninti, R. K.; Sengupta, A.; Hazra, P. Cucurbit[7]uril Assisted Ultraviolet to Visible Fluorescence Switch of a Heart Medicine. Phys. Chem. Chem. Phys. 2014, 16, 28232826. (26) Assaf, K. I.; Nau, W. M. Cucurbiturils: From Synthesis to High-Affinity Binding and Catalysis. Chem. Soc. Rev. 2015, 44, 394-418. (27) Fedorov, Y. V.; Tkachenko, S. V.; Chernikova, E. Y.; Godovikov, I. A.; Fedorova, O. A.; Isaacs, L. Photoinduced Guest Transformation Promotes Translocation of Guest from Hydroxypropyl-β-cyclodextrin to Cucurbit[7]uril. Chem. Commun. 2015, 51, 1349-1352. (28) Thomas, S. S.; Bohne, C. Determination of the Kinetics Underlying the pKa Shift for the 2-Aminoanthracenium Cation Binding with Cucurbit[7]uril. Faraday Discuss. 2015, 185, 381398. (29) Saleh, N. i.; Al-Handawi, M. B.; Bufaroosha, M. S.; Assaf, K. I.; Nau, W. M. Tuning Protonation States of Tripelennamine Antihistamines by Cucurbit[7]uril. J. Phys. Org. Chem. 2016, 29, 101-106. (30) He, S.; Sun, X.; Zhang, H. Influence of the Protonation State on the Binding Mode of Methyl Orange with Cucurbiturils. J. Mol. Struct. 2016, 1107, 182-188.

ACS Paragon Plus Environment

26

Page 27 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(31) Lee, J.; Miller, B. T.; Brooks, B. R. Computational Scheme for pH-dependent Binding Free Energy Calculation with Explicit Solvent. Protein Sci. 2016, 25, 231-243. (32) Sayed, M.; Pal, H. Supramolecularly Assisted Modulations in Chromophoric Properties and Their Possible Applications: An Overview. J. Mater. Chem. C 2016, 4, 2685-2706. (33) Liu, Y.; Guo, D.-S.; Zhang, H.-Y.; Ma, Y.-H.; Yang, E.-C. The Structure and Thermodynamics of Calix[n]arene Complexes with Dipyridines and Phenanthroline in Aqueous Solution Studied by Microcalorimetry and NMR Spectroscopy. J. Phys. Chem. B 2006, 110, 3428-3434. (34) Megyesi, M.; Biczók, L.; Jablonkai, I. Highly Sensitive Fluorescence Response to Inclusion Complex Formation of Berberine Alkaloid with Cucurbit[7]uril. J. Phys. Chem. C 2008, 112, 3410-3416. (35) Florea, M.; Nau, W. M. Implementation of Anion-Receptor Macrocycles in Supramolecular Tandem Assays for Enzymes Involving Nucleotides as Substrates, Products, and Cofactors. Org. Biomol. Chem. 2010, 8, 1033-1039. (36) Arimura, T.; Nagasaki, T.; Shinkai, S.; Matsuda, T. Host−Guest Properties of New Water-Soluble Calixarenes Derived from p-(Chloromethyl)calixarenes. J. Org. Chem. 1989, 54, 3766-3768. (37) Wagner,

B.

D.;

MacDonald,

P.

J.

The

Fluorescence

Enhancement

of

1-

Anilinonaphthalene-8-sulfonate (ANS) by Modified β-Cyclodextrins. J. Photochem. Photobiol., A 1998, 114, 151-157. (38) Wagner, B. D.; MacDonald, P. J.; Wagner, M. A Visual Demonstration of Supramolecular Chemistry: Observable Fluorescence Enhancement upon Host−Guest Inclusion. J. Chem. Educ. 2000, 77, 178-181.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 37

(39) Márquez, C.; Nau, W. M. Polarizabilities Inside Molecular Containers. Angew. Chem., Int. Ed. 2001, 40, 4387-4390. (40) Rankin, M. A.; Wagner, B. D. Fluorescence Enhancement of Curcumin upon Inclusion into Cucurbituril. Supramol. Chem. 2004, 16, 513-519. (41) Tuncel, D.; Özsar, Ö.; Tiftik, H. B.; Salih, B. Molecular Switch Based on a Cucurbit[6]uril Containing Bistable [3]Rotaxane. Chem. Commun. 2007, 1369-1371. (42) Bhasikuttan, A. C.; Choudhury, S. D.; Pal, H.; Mohanty, J. Supramolecular Assemblies of Thioflavin T with Cucurbiturils: Prospects of Cooperative and Competitive Metal Ion Binding. Isr. J. Chem. 2011, 51, 634-645. (43) Carvalho, C. P.; Uzunova, V. D.; Da Silva, J. P.; Nau, W. M.; Pischel, U. A Photoinduced pH Jump Applied to Drug Release from Cucurbit[7]uril. Chem. Commun. 2011, 47, 8793-8795. (44) de Silva, A. P.; Gunaratne, H. Q. N.; McCoy, C. P. A Molecular Photoionic AND Gate Based on Fluorescent Signaling. Nature 1993, 364, 42-44. (45) Ferreira, R.; Remón, P.; Pischel, U. Multivalued Logic with a Tristable Fluorescent Switch. J. Phys. Chem. C 2009, 113, 5805-5811. (46) Pischel, U.; Uzunova, V. D.; Remón, P.; Nau, W. M. Supramolecular Logic with Macrocyclic Input and Competitive Reset. Chem. Commun. 2010, 46, 2635-2637. (47) Ratner, T.; Keinan, E. Images Come to Life. ChemBioChem 2010, 11, 947-949. (48) Atilgan, S.; Akkaya, E. U. A Calixpyridinium−Pyranine Complex as a Selective Anion Sensing Assembly via the Indicator Displacement Strategy. Tetrahedron Lett. 2004, 45, 92699271.

ACS Paragon Plus Environment

28

Page 29 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(49) Zhu, L.; Anslyn, E. V. Facile Quantification of Enantiomeric Excess and Concentration with Indicator-Displacement Assays: An Example in the Analyses of α-Hydroxyacids. J. Am. Chem. Soc. 2004, 126, 3676-3677. (50) Bakirci, H.; Nau, W. M. Fluorescence Regeneration as a Signaling Principle for Choline and Carnitine Binding: A Refined Supramolecular Sensor System Based on a Fluorescent Azoalkane. Adv. Funct. Mater. 2006, 16, 237-242. (51) Nguyen, B. T.; Anslyn, E. V. Indicator–Displacement Assays. Coord. Chem. Rev. 2006, 250, 3118-3127. (52) Zhang, T.; Anslyn, E. V. Using an Indicator Displacement Assay to Monitor Glucose Oxidase Activity in Blood Serum. Org. Lett. 2007, 9, 1627-1629. (53) Ghale, G.; Nau, W. M. Dynamically Analyte-Responsive Macrocyclic Host−Fluorophore Systems. Acc. Chem. Res. 2014, 2150-2159. (54) Assaf, K. I.; Suckova, O.; Al Danaf, N.; von Glasenapp, V.; Gabel, D.; Nau, W. M. Dodecaborate-Functionalized Anchor Dyes for Cyclodextrin-Based Indicator Displacement Applications. Org. Lett. 2016, 18, 932-935. (55) Niikura, K.; Bisson, A. P.; Anslyn, E. V. Optical Sensing of Inorganic Anions Employing a Synthetic Receptor and Ionic Colorimetric Dyes. J. Chem. Soc., Perk. Trans. 2 1999, 11111114. (56) Nguyen, B. T.; Wiskur, S. L.; Anslyn, E. V. Using Indicator-Displacement Assays in Test Strips and To Follow Reaction Kinetics. Org. Lett. 2004, 6, 2499-2501. (57) Hennig, A.; Bakirci, H.; Nau, W. M. Label-Free Continuous Enzyme Assays with Macrocycle-Fluorescent Dye Complexes. Nat Methods 2007, 4, 629-632.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 37

(58) Bailey, D. M.; Hennig, A.; Uzunova, V. D.; Nau, W. M. Supramolecular Tandem Enzyme Assays for Multiparameter Sensor Arrays and Enantiomeric Excess Determination of Amino Acids. Chem. - Eur. J. 2008, 14, 6069-6077. (59) Nau, W. M.; Ghale, G.; Hennig, A.; Bakirci, H.; Bailey, D. M. Substrate-Selective Supramolecular Tandem Assays: Monitoring Enzyme Inhibition of Arginase and Diamine Oxidase by Fluorescent Dye Displacement from Calixarene and Cucurbituril Macrocycles. J. Am. Chem. Soc. 2009, 131, 11558-11570. (60) Florea, M.; Nau, W. M. Strong Binding of Hydrocarbons to Cucurbituril Probed by Fluorescent Dye Displacement: A Supramolecular Gas-Sensing Ensemble. Angew. Chem., Int. Ed. 2011, 50, 9338-9342. (61) Ghale, G.; Ramalingam, V.; Urbach, A. R.; Nau, W. M. Determining Protease Substrate Selectivity and Inhibition by Label-Free Suprannolecular Tandem Enzyme Assays. J. Am. Chem. Soc. 2011, 133, 7528-7535. (62) Guo, D.-S.; Uzunova, V. D.; Su, X.; Liu, Y.; Nau, W. M. Operational Calixarene-Based Fluorescent Sensing Systems for Choline and Acetylcholine and Their Application to Enzymatic Reactions. Chem. Sci. 2011, 2, 1722-1734. (63) Mock, W. L.; Shih, N. Y. Structure and Selectivity in Host−Guest Complexes of Cucurbituril. J. Org. Chem. 1986, 51, 4440-4446. (64) Mock, W. L. Cucurbituril; In Supramolecular Chemistry II — Host Design and Molecular Recognition; Weber, E., Ed. Springer: Berlin, Heidelberg, Germany, 1995; Vol. 175, pp. 1-24.

ACS Paragon Plus Environment

30

Page 31 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(65) Jon, S. Y.; Selvapalam, N.; Oh, D. H.; Kang, J.-K.; Kim, S.-Y.; Jeon, Y. J.; Lee, J. W.; Kim, K. Facile Synthesis of Cucurbit[n]uril Derivatives via Direct Functionalization: Expanding Utilization of Cucurbit[n]uril. J. Am. Chem. Soc. 2003, 125, 10186-10187. (66) Lagona, J.; Mukhopadhyay, P.; Chakrabarti, S.; Isaacs, L. The Cucurbit[n]uril Family. Angew. Chem., Int. Ed. 2005, 44, 4844-4870. (67) Liu, S.; Ruspic, C.; Mukhopadhyay, P.; Chakrabarti, S.; Zavalij, P. Y.; Isaacs, L. The Cucurbit[n]uril Family: Prime Components for Self-Sorting Systems. J. Am. Chem. Soc. 2005, 127, 15959-15967. (68) Kim, K.; Selvapalam, N.; Ko, Y. H.; Park, K. M.; Kim, D.; Kim, J. Functionalized Cucurbiturils and Their Applications. Chem. Soc. Rev. 2007, 36, 267-279. (69) Rekharsky, M. V.; Ko, Y. H.; Selvapalam, N.; Kim, K.; Inoue, Y. Complexation Thermodynamics of Cucurbit[6]uril with Aliphatic Alcohols, Amines, and Diamines. Supramol. Chem. 2007, 19, 39-46. (70) Klöck, C.; Dsouza, R. N.; Nau, W. M. Cucurbituril-Mediated Supramolecular Acid Catalysis. Org. Lett. 2009, 11, 2595-2598. (71) Wang, R.; Yuan, L.; Macartney, D. H. A Green to Blue Fluorescence Switch of Protonated 2-Aminoanthracene upon Inclusion in Cucurbit[7]uril. Chem. Commun. 2005, 58675869. (72) Gavvala, K.; Sengupta, A.; Koninti, R. K.; Hazra, P. Supramolecular Host-Inhibited Excited-State Proton Transfer and Fluorescence Switching of the Anti-Cancer Drug, Topotecan. ChemPhysChem 2013, 14, 3375-3383.

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 37

(73) Basílio, N.; Laia, C. A. T.; Pina, F. Excited-State Proton Transfer in Confined Medium. 4‑Methyl-7-hydroxyflavylium and β‑Naphthol Incorporated in Cucurbit[7]uril. J. Phys. Chem. B 2015, 119, 2749-2757. (74) Macartney, D. H. Cucurbit[n]uril Host-Guest Complexes of Acids, Photoacids, and Super Photoacids. Isr. J. Chem. 2017, 57, DOI: 10.1002/ijch.201700096. (75) Melo, M. J.; Melo, E.; Pina, F. Determination of Acid-Base Equilibria of Organic Pollutants: The Steady State Fluorescence Emission Method. Arch. Environ. Contam. Toxicol. 1994, 26, 510-520. (76) Frontier, A. J.; Raghavan, S.; Danishefsky, S. J. A Highly Stereoselective Total Synthesis of Hispidospermidin: Derivation of a Pharmacophore Model. J. Am. Chem. Soc. 2000, 122, 6151-6159. (77) Hamilton, R.; Walker, B. J.; Walker, B. A Convenient Synthesis of N-Protected Diphenyl Phosphonate Ester Analogs of Ornithine, Lysine and Homolysine. Tetrahedron Lett. 1993, 34, 2847-2850. (78) Shaikh, M.; Mohanty, J.; Singh, P. K.; Nau, W. M.; Pal, H. Complexation of Acridine Orange by Cucurbit[7]uril and β-Cyclodextrin: Photophysical Effects and pKa Shifts. Photochem. Photobiol. Sci. 2008, 7, 408-414. (79) Koner, A. L.; Ghosh, I.; Saleh, N. i.; Nau, W. M. Supramolecular Encapsulation of Benzimidazole-Derived Drugs by Cucurbit[7]uril. Can. J. Chem. 2011, 89, 139-147. (80) Shizuka, H.; Tsutsumi, K.; Takeuchi, H.; Tanaka, I. Proton Transfer Reactions in the Excited State of 1-Aminopyrene by Picosecond/Streak Camera and Nanosecond Spectroscopy. Chem. Phys. 1981, 59, 183-190.

ACS Paragon Plus Environment

32

Page 33 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(81) Bryson, A. The Effects of Substituents in the Naphthalene Ring − Part II. The Measurement of Second Dissociation Constants of the Aminonapthalene Monosulphonic Acids. Trans. Faraday Soc. 1951, 47, 522-527. (82) Hansen, J. E.; Pines, E.; Fleming, G. R. Excited-State Proton Transfer of Protonated 1Aminopyrene Complexed with β-Cyclodextrin. J. Phys. Chem. 1992, 96, 6904-6910. (83) Donckt, E. V.; Dramaix, R.; Nasielski, J.; Vogels, C. Photochemistry of Aromatic Compounds − Part 1. Acid-base Properties of Singlet and Triplet Excited States of Pyrene Derivatives and Aza-aromatic Compounds. Trans. Faraday Soc. 1969, 65, 3258-3262. (84) Rotkiewicz, K.; Grabowski, Z. R. Excited States of Aminoanthracenes. An Experimental Approach to Electron Density Distribution. Trans. Faraday Soc. 1969, 65, 3263-3278. (85) Hite, P.; Krasnansky, R.; Thomas, J. K. Spectroscopic Investigations of Surfaces Using Aminopyrene. J. Phys. Chem. 1986, 90, 5795-5799. (86) Tsutsumi, K.; Sekiguchi, S.; Shizuka, H. Proton-transfer Reactions in the Excited State of Phenanthrylamines by Nanosecond Spectroscopy and Fluorimetry. J. Chem. Soc., Faraday Trans. 1 1982, 78, 1087-1101. (87) Bakirci, H.; Koner, A. L.; Schwarzlose, T.; Nau, W. M. Analysis of Host-Assisted Guest Protonation Exemplified for p-Sulfonatocalix[4]arene—Towards Enzyme-Mimetic pKa Shifts. Chem. - Eur. J. 2006, 12, 4799-4807. (88) Pluth, M. D.; Bergman, R. G.; Raymond, K. N. Acid Catalysis in Basic Solution: A Supramolecular Host Promotes Orthoformate Hydrolysis. Science 2007, 316, 85-88. (89) Pluth, M. D.; Bergman, R. G.; Raymond, K. N. Making Amines Strong Bases: Thermodynamic Stabilization of Protonated Guests in a Highly-Charged Supramolecular Host. J. Am. Chem. Soc. 2007, 129, 11459-11467.

ACS Paragon Plus Environment

33

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 37

(90) Sebők-Nagy, K.; Biczók, L. Effect of Protonation and Hydrogen Bonding on the Fluorescent Properties and Exciplex Formation of N-(4-pyridyl)-1,2-naphthalimide. Photochem. Photobiol. Sci. 2004, 3, 389-395. (91) Koner, A. L.; Schatz, J.; Nau, W. M.; Pischel, U. Selective Sensing of Citrate by a Supramolecular 1,8-Naphthalimide/Calix[4]arene Assembly via Complexation-Modulated pKa Shifts in a Ternary Complex. J. Org. Chem. 2007, 72, 3889-3895. (92) Macartney, D. H. Encapsulation of Drug Molecules by Cucurbiturils: Effects on their Chemical Properties in Aqueous Solution. Isr. J. Chem. 2011, 51, 600-615. (93) Saleh, N. i.; Al-Soud, Y. A.; Al-Kaabi, L.; Ghosh, I.; Nau, W. M. A Coumarin-Based Fluorescent PET Sensor Utilizing Supramolecular pKa Shifts. Tetrahedron Lett. 2011, 52, 52495254. (94) Basílio, N.; Gago, S.; Parola, A. J.; Pina, F. Contrasting pKa Shifts in Cucurbit[7]uril Host–Guest Complexes Governed by an Interplay of Hydrophobic Effects and Electrostatic Interactions. ACS Omega 2017, 2, 70-75. (95) Dsouza, R. N.; Pischel, U.; Nau, W. M. Fluorescent Dyes and Their Supramolecular Host/Guest Complexes with Macrocycles in Aqueous Solution. Chem. Rev. 2011, 111, 79417980. (96) Datta, S.; Halder, M. Effect of Encapsulation in the Anion Receptor Pocket of Subdomain IIA of Human Serum Albumin on the Modulation of pKa of Warfarin and Structurally Similar Acidic Guests: A Possible Implication on Biological Activity. J. Photochem. Photobiol., B 2014, 130, 76-85. (97) Honig, B.; Nicholls, A. Classical Electrostatics in Biology and Chemistry. Science 1995, 268, 1144-1149.

ACS Paragon Plus Environment

34

Page 35 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(98) MacKerell, A. D., Jr.; Sommer, M. S.; Karplus, M. pH Dependence of Binding Reactions from Free Energy Simulations and Macroscopic Continuum Electrostatic Calculations: Application to 2'GMP/3'GMP inding to Ribonuclease T1 and Implications for Catalysis. J. Mol. Biol. 1995, 247, 774-807. (99) Ha, N.-C.; Kim, M.-S.; Lee, W.; Choi, K. Y.; Oh, B.-H. Detection of Large pKa Perturbations of an Inhibitor and a Catalytic Group at an Enzyme Active Site, a Mechanistic Basis for Catalytic Power of Many Enzymes. J. Biol. Chem. 2000, 275, 41100-41106. (100) Matulis, D.; Bloomfield, V. A. Thermodynamics of the Hydrophobic Effect. I. Coupling of Aggregation and pKa Shifts in Solutions of Aliphatic Amines. Biophys. Chem. 2001, 93, 3751. (101) Popov, M. E.; Sten’gach, M. A.; Andreeva, N. S. Modeling of Substrate and Inhibitory Complexes of Histidine–Aspartic Protease. Russ. J. Bioorg. Chem. 2008, 34, 380-386. (102) Márquez, C.; Nau, W. M. Two Mechanisms of Slow Host−Guest Complexation between Cucurbit[6]uril and Cyclohexylmethylamine: pH-Responsive Supramolecular Kinetics. Angew. Chem., Int. Ed. 2001, 40, 3155-3160. (103) Bakirci, H.; Koner, A. L.; Nau, W. M. Spherical Shape Complementarity as an Overriding Motif in the Molecular Recognition of Noncharged Organic Guests by pSulfonatocalix[4]arene: Complexation of Bicyclic Azoalkanes. J. Org. Chem. 2005, 70, 99609966. (104) Liu, Y.; Li, C.-J.; Guo, D.-S.; Pan, Z.-H.; Li, Z. A Comparative Study of Complexation of β-Cyclodextrin, Calix[4]arenesulfonate and Cucurbit[7]uril with Dye Guests: Fluorescence Behavior and Binding Ability. Supramol. Chem. 2007, 19, 517-523.

ACS Paragon Plus Environment

35

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 37

(105) Assaf, K. I.; Nau, W. M. Cucurbiturils as Fluorophilic Receptors. Supramol. Chem. 2014, 26, 657-669. (106) Gamal-Eldin, M. A.; Macartney, D. H. Cucurbit[7]uril Host−Guest Complexations of Steroidal Neuromuscular Blocking Agents in Aqueous Solution. Can. J. Chem. 2014, 92, 243249. (107) Basílio, N.; Mendoza, J.; Gago, S.; Parola, A. J. pH-Driven Self-Sorting in a Four Component Host–Guest System. Chem. Commun. 2017, 53, 6472-6475. (108) Dsouza, R. N.; Hennig, A.; Nau, W. M. Supramolecular Tandem Enzyme Assays. Chem. - Eur. J. 2012, 18, 3444-3459.

ACS Paragon Plus Environment

36

Page 37 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table of Contents Graphic

ACS Paragon Plus Environment

37