Abstract
The pyronin class of fluorophores serves a critical role in numerous imaging applications, particularly involving preferential staining of RNA through base pair intercalation. Despite this important role in molecular staining applications, the same set of century-old pyronins (i.e. pyronin Y (PY) and pyronin B (PB)), which possess relatively low fluorophore brightness, are still predominantly being used due to the lack of methodology for generating enhanced variants. Here, we use TD-DFT calculations of interconversion energies between structures on the S1 surface as a preliminary means to evaluate fluorophore brightness for a proposed set of pyronins containing variable substitution patterns at the 2, 3, 6, and 7 positions. Using a nucleophilic aromatic substitution/hydride addition approach, we synthesized the same set of pyronins, and demonstrate that quantum-mechanical computations are useful for predicting fluorophore performance. We produced the brightest series of pyronin fluorophores described to date, which possess considerable gains over PY and PB.
Xanthene-based structures with 3,6-diamino substitution, more commonly referred to as pyronins (Figure 1),1 have origins in the late 19th century2, 3 and became popular as dyes originally in textiles then later in the scientific community for imaging applications.4, 5 Pyronins, like structurally similar rhodamines, typically exhibit fluorescent properties and are commonly employed in microscopy,6, 7 cell cycle studies,8 electrophoresis,9, 10 flow cytometry,11, 12 paraffin section analysis,13, 14 mitochondrial studies,15–17 detecting protein-protein interactions,18 chemo/ion sensors,16, 19 among other uses.20–24 Most applications utilize commercially available pyronin Y (PY) or pyronin B (PB) and to a lesser extent acridine red (AR) which preferentially intercalate RNA allowing for imaging/detection at an excitation wavelength (λex) of ~530 nm (AR)/~550 nm (PY/PB) and an emission wavelength (λem) of ~560 nm (AR)/~570 nm (PY/PB), however, numerous synthetically tailored pyronins have been generated for specialized uses.15, 18, 21 While PB, AR, and particularly PY, remain popular dyes, the quantum yields (Φf) for these fluorophores in aqueous solvents are relatively low in comparison to other xanthene-based fluorophores such as rhodamines and fluoresceins limiting the sensitivity of these fluorophore systems. As a result, pyronins are often applied in high concentrations (~100 μM) to stain cells which can lead to cytotoxicity.
Figure 1.
Structures and quantum yields1 of alkyl substituted pyronins.
In general, fluorophores containing arylamino groups, like that of PB, PY, and AR, are prone to undergo twisted intramolecular charge transfer (TICT) mechanisms in the excited singlet state which can reduce the quantum efficiency of the system (Figure 2).25 In this process, alkyl substituted arylamino groups rotate out of conjugation with the π-system resulting in an electron transfer that produces a diradical form of the molecule. Recently, efforts have been made to mitigate TICT mechanisms as a means to produce fluorophores that possess particularly high quantum yields and molar absorptivities.26 This concept has been applied by Lavis et al,27, 28 in which N,N-dialkylamino groups were replaced with azetidinyl substitution to dramatically increase the quantum efficiency and photon yield for a range of classical fluorophores; N-arylazetidine systems possess larger energetic barriers for electron transfer processes needed to form the TICT state in comparison to N,N-aryldialkylamino systems.28
Figure 2.
Jabłoński diagram illustrating an excited singlet state twisted intramolecular charge transfer (TICT).
We became interested in optimizing the photophysical properties of pyronins, to generate more photogenic variants that could be used in considerably lower cellular staining concentrations and/or could provide higher resolution and more information rich data. Our envisioned approach, was the replacement of traditional N,N-substituted amino groups at the 3 and 6 position with groups that produce relatively large energetic barriers for TICT processes in the excited state, in order to maximize molar absorptivities and quantum yield efficiencies. We were particularly interested in the replacement of dimethylamino (PY)/diethylamino (PB) groups with an azetidinyl group, due to gains observed in structurally similar rhodamine-based systems,28 as well as monoalkylamino substitution patterns due to the >2-fold increase in quantum yield between PY/PB and AR.1 Additionally, we were interested in introducing 2,7-difluoro-substitution into the pyronin ring structure, for some variants, to attenuate hydrophobicity and investigate the potential of NH-F intramolecular hydrogen bonding as a means deter TICT processes; analogous fluorination substitution has been successfully incorporated into fluorescein (i.e. Oregon Green fluorophore)29 and rhodamine derivatives30 to improve photophysical, chemical, and biological properties while preserving fluorophore brightness. Furthermore, we aimed to retain the relative polarity properties of PY, PB, and AR, as these pyronins possess generally good aqueous solubility and cellular permeability which is critical for effective cellular staining. To aid our design, we performed Log P and quantum-mechanical time-dependent density functional theory (TD-DFT) calculations as well as calculated vibronically-resolved fluorescence spectra to evaluate the lipophilicity and propensity to undergo excited singlet state TICT for a range of 3,6-diamino substituted pyronins.
We envisioned computationally evaluating a set of pyronins (1-13, Figure 3) containing various amino-substitution patterns at the 3,6-positions, in addition to the incorporation of hydrogen (1-5) or fluorine (6-12) substituents at the 2,7-positions; PY, PB, and AR were used for comparison. The proposed series contains azetidine-substituted pyronins 1 and 8, monoalkylamino members variants of AR (2-5), and 2,7-difluorinated analogs including fluorinated versions of PY (7), PB (8), and AR (9) along with 10–13 which are capable of forming H-F intramolecular hydrogen bonds. To evaluate the hydrophobicity, calculated Log P (cLog P) values were generated31 for 1-13, PY, PB, and AR (Table 1). To ensure retention of the relative lipophilicity of PY, PB, and AR a cutoff of <1 cLog P was employed for 1-13 removing butylamino substituted pyronins 5 and 13 from further analysis.
Figure 3.
Structures of pyronins 1-13 and PY, PB, and AR with calculated Log P (cLog P) values31 (in blue) with H-F hydrogen bonding illustrated for 9-13 (red dashes).
Table 1.
Energy differences and interconversion barriers (kcal/mol) between S1-min and TICT states of pyronins 1-13, with PY, PB, and AR from TD-DFT calculations. Dihedral angles (DA) between the aromatic plane and alkylamino groups are given in degrees (°) for TICT structures in the first excited state.
| ΔG | ΔG‡ | DA | ΔG | ΔG‡ | DA | ||
|---|---|---|---|---|---|---|---|
|
| |||||||
| PY | −3.2 | 5.1 | 89.6 | 6 | −8.2 | 1.0 | 90.4 |
| PB | −4.2 | 3.1 | 88.5 | 7 | −6.6 | 1.4 | 89.6 |
| AR | +0.1 | 8.8 | 0.2 | 8 | +0.3 | 7.1 | 89.7 |
| 1 | +2.4 | 9.1 | 89.9 | 9 | +2.8 | 8.5 | 1.4 |
| 2 | −0.0 | 7.2 | 0.1 | 10 | +3.0 | 7.7 | 2.3 |
| 3 | −0.1 | 8.4 | 0.1 | 11 | +3.0 | 8.5 | 7.1 |
| 4 | −0.2 | 7.4 | 2.2 | 12 | +3.3 | 7.3 | 2.3 |
For the remaining pyronins, geometry optimizations were performed with the B3LYP-D3BJ DFT32–35 functional and def2-TZVP36 basis sets while solvation effects in water were described with CPCM37–39 using ORCA 5.0.2 package.40–42 The locally-excited structures, labelled as S1-min, as well as the TICT structures were optimized with Tamm-Dancoff approximation43 TD-DFT on the first excited state surface. The S1-min structures retain the structural symmetry of the S0 lowest-energy structures, in other words, the side groups have significant overlap in both states. Vibrational frequency analyses were performed for all structures to determine their local minima or saddle point characters. Gibbs free energy differences, ΔG, between S1-min and TICT structures as well as interconversion barriers, ΔG‡, were determined (Table 1). A positive ΔG value shows the S1-min structure is lower in energy than the TICT structure. By contrast, a negative ΔG value shows the TICT structure is lower in energy. We emphasize that the ΔG‡ barriers arise from rotation of the alkylamino side chains (Table 1). Transition states between the S1-min and TICT structures were obtained with climbing-image nudged-elastic band, CI-NEB, calculations on the first excited-state surface.
On the first excited state surface, the arylamino alkyl groups are fully rotated between the xanthene core for 2-4 and 9-12 (monoalkylamino species), restoring near-planarity (Table 1 and Figure S38). By contrast, for the dialkylamino species, the TICT structures represent partial (approximately 90°) rotation of the arylamino alkyl groups (Table 1). S1-min and TICT structures of the monoalkylamino species (AR, 2, 3 and 4) are isoenergetic. The rigid azetidinyl group (1) favors S1-min over the TICT structure by 2.4 kcal/mol, deterring S1-min/TICT interconversion. Likewise, 2,7-difluoro groups (9, 10, 11 and 12) favor S1-min geometries by 2.8–3.3 kcal/mol. Interestingly, when the azetidinyl and 2,7-difluoro strategies are combined (8), the TICT structure is isoenergetic with S1-min. ΔG for 8 is only +0.3 kcal/mol, compared with +2.4 kcal/mol for 1.
Comparing PY and PB respectively to 6 and 7, shows impact of 2,7-difluoro substitutions for dialkylamino species. For 6 and 7, TICT structures are lower in energy than S1-min geometries by 6.6–8.2 kcal/mol. Without fluorine substitution, the S1-min structures of PY and PB are stabilized (−3.2 to −4.2 kcal/mol, Table 1), although TICT structures remain lower in energy. Overall, Table 1 shows that the S1-min → TICT interconversion free energies are increased by 2,7-difluoro substitution in monoalkylamino species (AR, 2, 3 and 4 versus 9, 10, 11 and 12). The opposite is true for the dialkylamino species (1, 6 and 7 versus 8, PY and PB). Thus, fluorination is only useful for deterring S1-min → TICT interconversion in the monoalkylamino species. Based on ΔG values in Table 1, contributions of TICT geometries to the population on the first-excited state surface follow the trend: 6 and 7>> PY and PB >> AR, 2-4 and 8 > 1 and 9–12. Thus, TICT geometries are least accessible for 1 and 9-12.
Comparison of the vertical excitation energies of 10 and 7 were carried out (see Figure S38 for a visualization of this comparison). These are used as examples as they belong to groups on opposite extremes of the data in Table 1. For 10, the first absorption, S0 → S1, is a π-π* excitation dominated by the configuration of HOMO/LUMO transition, with both orbitals centered at the xanthene core. This is the case for the S0 minimum structure as well as a structure with a twisted side-group. For 7, rotation of a diethylamino side group by about 90° leads to a twisted structure similar to the TICT geometry on the first excited state surface. Crucially, the HOMO is now centered at the amino group although the LUMO remains centered at the xanthene core. Natural transition orbitals are provided in Figure S38. Thus, the HOMO-LUMO transition in this twisted structure of 7 leads to a dark state (Figure 4). The low oscillator strength for the TICT-like geometry of 7 reveals that it can only relax via non-radiative decay. This is also the case for 6 and is fully consistent with earlier reports on rhodamine dyes by Rega et al.44 who reported that structural deformation of the xanthene and carboxyphenyl groups can switch the energy orderings of two excited states with dark and bright characters. We conclude by emphasizing that Table 1 indicates that non-radiative decay will be highest for PY, PB, 6 and 7 for which the TICT structures are lower in energy on the first excited state surface. Conversely, we expect improved quantum yields for AR, 1-4, and 8-12.
Figure 4.
Optimized ground state geometry of 7 with a twisted diethylamino side group. Frontier orbitals for this structure are also shown.
Encouraged by our computational results, we pursued the synthesis of 1-4 and 6-12 through a two-staged synthetic route (Figure 5), initiating from hexafluorobenzophenone 14 or difluoroxanthone 15. The first stage utilizes a nucleophilic aromatic substitution (SNAr) approach to substitute fluorine atoms, situated para to the carbonyl group of 1445 or 15,46 with amino groups; additional hydroxyl substitution of 14 at an ortho substituted fluorine provides the xanthone ring structure via intramolecular cyclization. The second stage uses hydride addition to the carbonyl, which subsequently undergoes elimination to provide the pyronin product.
Figure 5.
SNAr/hydride addition approach to generating pyronins.
First stage reaction conditions for the substitution of 14 gave generally high reaction yields for the two-step sequence involving the treatment with amine followed by heating in hydroxide to provide the xanthone (Table 2). While most reaction yields for this conversion were >87%, the conversion of 14 to 19 underwent substantial N-demethylation during treatment with hydroxide. As a consequence of this low-yielding step, we decided to remove pyronin 9 from our synthetic library. Amino substitution conditions were applied to xanthone 15, with heating, which provided 3,6-diamine substituted products 23-26 in good to excellent yields (Table 3).
Table 2.
Synthesis of xanthones 16-22 using SNAr conditions.
| |||
|---|---|---|---|
|
| |||
| Entry | HNR2a | Yield (%)d | Product |
|
| |||
| 1 | HNMe2b | 94 | 16 |
| 2 | HNEt2 | 87 | 17 |
| 3 | azetidinec | 84 | 18 |
| 4 | H2NMe | <5e | 19 |
| 5 | H2NEt | 92 | 20 |
| 6 | H2NPr | 91 | 21 |
| 7 | H2NiPr | 94 | 22 |
All amines were used neat and were reacted at 26 °C with the exception of HNMe2 and azetidine.
Dimethylamine was generated via decomposition of DMF in KOH at 60 °C.
Azetidine was reacted as a solution in acetonitrile at 60 °C.
Isolated yield.
Product of the second step decomposes in presence of hot KOH.
Table 3.
Synthesis of xanthones 23-26 using SNAr conditions.
All amines were used as aqueous mixtures at 200 °C in a sealed-vessel with the exception of azetidine.
Azetidine was reacted in a sealed-vessel as a solution in acetonitrile at 150 °C.
Isolated yield.
Second stage reaction conditions, involved the reduction of xanthones 16-18 and 20-26 using borane-dimethylsulfide complex (BH3-DMS). Similarly to previously reported reductions of this type,32 two equivalents of hydride are added to the 9-position to provide the over-reduced xanthene product, which slowly decomposes upon standing, in the presence of atmospheric oxygen, to the pyronin product. This oxidation process is expediated in the presence of 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), to provide pyronin product as a trifluoroacetic acid salt, upon acidic workup. Table 4 summarizes the yield for this two-step reduction/oxidation process, which produces pyronin products 1-4, 6-8, and 10-12 in good to excellent yields.
Table 4.
Synthesis of pyronins 1-4, 6-8, 10-12 and comparison of photophysical propertiesa with PY, PB, and AR.
| ||||||
|---|---|---|---|---|---|---|
|
| ||||||
| Yield (%)b | λmax (nm) | εmax (cm−1M−1) |
λem (nm) | Φf | εmax X Φf | |
|
| ||||||
| PY | - | 545 | 42,600 | 563 | 0.18 | 7,700 |
| PB | - | 551 | 65,600 | 568 | 0.20 | 13,100 |
| AR | - | 519 | 44,900 | 535 | 0.62 | 26,000 |
| 1 | 82 (23) | 545 | 70,600 | 563 | 0.72 | 50,800 |
| 2 | 78 (24) | 519 | 66,000 | 537 | 0.81 | 53,500 |
| 3 | 84 (25) | 522 | 66,800 | 539 | 0.82 | 54,800 |
| 4 | 86 (26) | 522 | 43,500 | 539 | 0.87 | 37,800 |
| 6 | 90 (16) | 554 | 34,400 | 573 | 0.01 | 3,400 |
| 7 | 91 (17) | 561 | 61,400 | 578 | 0.02 | 12,200 |
| 8 | 86 (18) | 549 | 71,700 | 567 | 0.63 | 45,200 |
| 10 | 75 (20) | 516 | 57,800 | 534 | 0.76 | 43,900 |
| 11 | 77 (21) | 519 | 56,200 | 536 | 0.93 | 52,300 |
| 12 | 75 (22) | 520 | 49,200 | 536 | 0.75 | 36,900 |
Photophysical properties were measured in pH 7.0 PBS buffer.
Isolated yield from reactant (in parenthesis).
The photophysical properties of pyronins 1-4, 6-8, and 10-12 were acquired and compared to PY, PB, and AR as summarized in Table 4 (see Figure S1 for experimental spectra). The maximum absorbance values (λmax) for dialkylamino-substituted pyronins 1, and 6-8 are analogous to PY and PB and structurally similar rhodamines (i.e. rhodamine B) and can be excited with lasers in the 550 nm range.
Vibronically-resolved absorption and fluorescence spectra of 1-4, 8, 10-12 and AR were computed with the adiabatic Hessian (AH) approach, with the ESD module of ORCA version 5.0.2. Details are provided in the Supporting Information, SI.47.48–51 To test dependence of our results on the DFT functional, we used ωB97X-D3,52, 53 PBE-D3BJ33, 34, 54 and BLYP-D3BJ.33–35, 55 We also considered dependence of the calculated spectra on the percentage of Hartree-Fock (HF) exchange in B3LYP, specifically we tested 0.05 (B3LYP-5%), 0.10 (B3LYP-10%) or 0.15 (B3LYP-15%). Vibronic spectra are compared to experimental data and vertical excitations obtained with the STEOM-DLPNO-CCSD approach in the SI.48, 56 We did not compute the vibronic spectra for PY, PB, 6 and 7. This is because our analysis that TICT structures are lower in energy than S1-min geometries on the first excited state surface (Table 1). Additionally, the nature of the S0 → S1 transition is significantly different for TICT structures of 6 and 7 (Figure 4). Treatment of these systems will likely require multiconfigurational treatment of the excited state wavefunctions as well as methodologies to account for the structural relaxations on the excited state surface, far beyond the scope of the current work.
Comparing absorption band maxima, λmax, obtained from TD-DFT spectra to experimental data (Table 4) shows that B3LYP-D3BJ consistently underestimates the experimental wavelength by 27–33 nm (Table 5). ωB97X-D3 shifts to even shorter wavelengths, by as much as 72–96 nm, while PBE-D3BJ overestimates the experiment to longer wavelengths, by 16–40 nm. Better agreement is obtained by reducing the amount of Hartree-Fock exchange in the B3LYP functional, to either 5% or 10%.,
Table 5.
Deviation of absorption and fluorescence λmax (nm) from TD-DFT from experimental data.
| Absorption | Fluorescence | |||
|---|---|---|---|---|
|
| ||||
| MSD | MAD | MAD | MSD | |
|
| ||||
| PBE-D3BJ | +23 | 23 | +11 | 11 |
| BLYP-D3BJ | +25 | 25 | +12 | 12 |
| B3LYP-5% | +5 | 5 | −10 | 10 |
| B3LYP-10% | −10 | 10 | −24 | 24 |
| B3LYP-15% | −22 | 22 | −37 | 37 |
| B3LYP-D3BJ | −31 | 31 | −46 | 46 |
| ωB97X-D3 | −83 | 83 | −94 | 94 |
| ωB97X | −85 | 85 | −99 | 99 |
| CAM-B3LYP | −71 | 71 | −85 | 85 |
Experimentally, monoalkyl-substituted pyronins 2-4 and 10-12 possessed λmax values of approximately 525 nm, and demonstrate similar bathochromic shifts to AR in comparison to PY/PB. The TD-DFT absorption data indeed shows that the band maxima for 2-4 and 10-12 are indeed within 5 nm of AR’s band maximum, with all DFT functionals (see Figure S39). Experiments also indicate that the band maxima of the azetidinyl compounds, 1 and 8, are bathochromically shifted by 26–30 nm versus AR. All hybrid functionals replicate these bathochromic shifts, 31–44 nm (Figure S39).
For fluorescence band maxima, B3LYP-D3BJ and ωB97X-D3 consistently underestimate experimental wavelengths (Table 5 and Figure S40). The bathochromic shifts of 1 and 8, versus AR, 2-4 and 9-12 are however captured by both functionals, although the lowest MADs are obtained for B3LYP-5% and BLYP-D3BJ.57
Pyronins 1-4, 6-8, and 10-12 all possess Stoke shift values in the range of 16–19 nm, again baring strong resemblance to the photophysical trends observed for PY, PB, and AR. However, in most cases, the molar absorptivity at λmax (εmax) and the quantum yields (Φf) are elevated in comparison to PY, PB, and AR, indicating that these pyronin variants, as was predicted from our computational studies, are brighter fluorophores than PY, PB, and AR. Notice that pyronins 6 and 7, have very low Φf values, due to a more stabilized TICT structure in the excited state, as described earlier in our computational studies, Table 1. In general, experimental Φf values are in moderately decent alignment with our predictions (based on the integrated vibronic fluorescence spectra, see SI), with 11 > 4 > 2 ~ 3 > 10 ~ 12 > 1 > 8.
In considering εmax x Φf, a quantitative measure of fluorophore intensity, other trends emerge. Most notably, n-propylamino substitution results in highly fluorescent pyronins 3 and 11, though unexpectantly pyronin 3 demonstrated slightly brighter fluorescence than 11. Monoalkyl-substituted pyronins 2-4, in general, result in strong fluorophore performance, and all out-perform fluorinated counterparts (10-12). Interestingly, isopropylamino substitution lowers molar absorptivity which adversely affects fluorophore brightness, despite relatively high quantum yields. Azetidinyl substitution, as noted in previous studies,24, 58 produced highly fluorescent pyronins in the 550 nm excitation range; nevertheless, both 1 and 8 are less fluorescent than monoalkylamino species 2, 3, and 11, which appear to better mitigate TICT transitions states due to larger energetic barriers of rotation about the amino group in the S1-min structures. Overall, fluorophore intensities follow the trend 3 > 2 > 11 > 1 > 8 > 10 > 4 > 12 >> PB > PY.
In summary, using molecular modeled interconversion energies between S1 and TICT states and calculated vibronic adsorption and fluorescence spectra, we were able to predict large differences in fluorophore performance for the pyronin family of fluorophores. In particular, the S1 ↔ TICT interconversion energies are very useful for separating species that either do or do not possess lossy internal conversion to the dark TICT states. The integrated areas for calculated vibronic fluorescence spectra and absorption oscillator strengths are also useful for estimating quantum yield trends, which often correlates to overall fluorophore brightness. Methodologically, we found that while TD-DFT generally reproduces the trend in absorption and fluorescence band maxima, the best agreement, for these set of compounds, is found with DFT functionals with small amounts of exact Hartree-Fock exchange, 0–10%. Through the computational methods adopted in this work, we were able to identify the 3,6-dipropylamino substitution patterns as a means for generating highly fluorescent pyronins 3 and 11 in the 525 nm excitation range. Additionally, 3,6-diazetinyl substitution of 1 and 8 offer bright pyronin fluorophores in the 550 nm excitation range. Together, pyronins 1, 3, 8, and 11 offer considerable gains (~2–6.5 fold increases in εmax x Φf) over traditional pyronins (i.e. AR, PY, and PB), and are the most intensely fluorescent set of pyronin fluorophores reported to date. It is likely that information obtained from molecular modeling of energetic barriers between S1 and TICT states could be used as a means to guide the discovery of other high performance xanthene-based fluorophores such as within the acridine, rhodamine, and rhodols families.
Supplementary Material
ACKNOWLEDGMENT
Z.R.W., B.W., A.H., C.A., and F.B. thank the NIH (2P20 GM103440-14A1) and NSF EPSCoR (IIA-1301726) for their generous funding.
Funding Sources
NIH (2P20 GM103440-14A1) and NSF EPSCoR (IIA-1301726)
ABBREVIATIONS
- PY
pyronin Y
- PB
pyronin B
- AR
acridine red
- TICT
twisted intramolecular charge transfer
- TD-DFT
quantum-mechanical time-dependent density functional theory
- cLog P
calculated Log partition coefficient
- ΔE
energy differences between S1 and TICT structures
- ΔE‡
interconversion energy barriers between S1 and TICT structures
- ΔH
enthalpy differences between S1 and TICT structures
- ΔH‡
interconversion enthalpy barriers between S1 and TICT structures
- HOMO
highest occupied molecular orbital
- LUMO
lowest unoccupied molecular orbital
- SNAr
nucleophilic aromatic substitution
- DDQ
2,3-dichloro-5,6-dicyano-1,4-benzoquinone
- λmax
maximum absorbance wavelength
- εmax
molar absorptivity at the maximum absorbance value
- Φf
quantum yield
- λem
maximum emission wavelength
Footnotes
ASSOCIATED CONTENT
The Supporting Information is available free of charge on the ACS Publications website. 1H NMR, 13C NMR, molar absorptivity plots and quantum yield plots (PDF).
The authors declare no competing financial interests.
REFERENCES
- (1).Zhang X-F; Zhang J; Lu X. The Fluorescence Properties of Three Rhodamine Dye Analogues: Acridine Red, Pyronin Y and Pyronin B. J. Fluoresc 2015, 25, 1151–1158. [DOI] [PubMed] [Google Scholar]
- (2).Biehringer J. Pyronines. Ber. 1894, 27, 3299. [Google Scholar]
- (3).Biehringer J. Dyes of the pyronine group. J. pr. Chem 1896, 54, 217. [Google Scholar]
- (4).Anhaeusser H. Determination of the authenticity of rapeseeds. Angew. Bot 1954, 28, 13–40. [Google Scholar]
- (5).Mitsumura H. Influence of orotic acid on Ehrlich ascites tumor cells. Bitamin 1961, 22, 336. [Google Scholar]
- (6).Martins JR; Bitondi MMG The HEX 110 hexamerin is a cytoplasmic and nucleolar protein in the ovaries of Apis mellifera. PLoS One 2016, 11, e0151035/0151031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Miao Y; Qian N; Shi L; Hu F; Min W. 9-Cyanopyronin probe palette for super-multiplexed vibrational imaging. Nat. Commun 2021, 12, 4518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Crissman HA; Darzynkieicz Z; Tobey RA; Steinkamp JA Correlated measurements of DNA, RNA, and protein in individual cells by flow cytometry. Science 1985, 228, 1321–1324. [DOI] [PubMed] [Google Scholar]
- (9).Nikolov TK; Uriel J; Christol G; Grabar P. Studies by agar-gel electrophoresis of the nucleoproteins and nucleic acids in tissue extracts. Rev. Fr. Etud. Clin. Biol 1962, 7, 519. [PubMed] [Google Scholar]
- (10).Bachvaroff R; McMaster PRB Separation of microsomal RNA into five bands during agar electrophoresis. Science 1964, 143, 1177–1179. [DOI] [PubMed] [Google Scholar]
- (11).Tanke HJ; Nieuwenhuis IA; Koper GJ; Slats JC; Ploem JS Flow cytometry of human reticulocytes based on RNA fluorescence. Cytometry 1981, 1 , 313–320. [DOI] [PubMed] [Google Scholar]
- (12).Shapiro HM Flow cytometric estimation of DNA and RNA content in intact cells stained with Hoechst 33342 and pyronin Y. Cytometry 1981, 2, 143–150. [DOI] [PubMed] [Google Scholar]
- (13).Chen CM; Chang HL Methyl green-pyronin as a stain for mast cells in paraffin sections. Stain Technol. 1963, 38, 133–135. [PubMed] [Google Scholar]
- (14).Li B; Wu Y; Gao X-M Pyronin Y as a fluorescent stain for paraffin sections. Histochem. J 2002, 34, 299–303. [DOI] [PubMed] [Google Scholar]
- (15).Morozumi A; Kamiya M; Uno S -n.; Umezawa, K.; Kojima, R.; Yoshihara, T.; Tobita, S.; Urano, Y. Spontaneously Blinking Fluorophores Based on Nucleophilic Addition/Dissociation of Intracellular Glutathione for Live-Cell Super-resolution Imaging. J. Am. Chem. Soc 2020, 142, 9625–9633. [DOI] [PubMed] [Google Scholar]
- (16).Yang M; Fan J; Sun W; Du J; Peng X. Mitochondria-Anchored Colorimetric and Ratiometric Fluorescent Chemosensor for Visualizing Cysteine/Homocysteine in Living Cells and Daphnia magna Model. Anal. Chem 2019, 91, 12531–12537. [DOI] [PubMed] [Google Scholar]
- (17).Sun Y-Q; Liu J; Zhang H; Huo Y; Lv X; Shi Y; Guo W. A Mitochondria-Targetable Fluorescent Probe for Dual-Channel NO Imaging Assisted by Intracellular Cysteine and Glutathione. J. Am. Chem. Soc 2014, 136, 12520–12523. [DOI] [PubMed] [Google Scholar]
- (18).Hymel D; Woydziak ZR; Peterson BR Detection of Protein–Protein Interactions by Proximity-Driven SNAr Reactions of Lysine-Linked Fluorophores. J. Am. Chem. Soc 2014, 136, 5241–5244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Kurishita Y; Kohira T; Ojida A; Hamachi I. Organelle-Localizable Fluorescent Chemosensors for Site-Specific Multicolor Imaging of Nucleoside Polyphosphate Dynamics in Living Cells. J. Am. Chem. Soc 2012, 134, 18779–18789. [DOI] [PubMed] [Google Scholar]
- (20).Fan J; Wu E; Dong J; Zhu R; Li M; Gao J; Han H; Ding L. A minimalist ratiometric fluorescent sensor based on non-covalent ternary platform for sensing H2S in aqueous solution and serum. Colloids Surf., A 2021, 616, 126299. [Google Scholar]
- (21).Song Y; Zhang H; Wang X; Geng X; Sun Y; Liu J; Li Z. One Stone, Three Birds: pH Triggered Transformation of Aminopyronine and Iminopyronine Based Lysosome Targeting Viscosity Probe for Cancer Visualization. Anal. Chem 2021, 93, 1786–1791. [DOI] [PubMed] [Google Scholar]
- (22).Kim DH; Park NH; Kim TW Highly efficient flexible organic light-emitting devices based on PEDOT:PSS electrodes doped with highly conductive Pyronin B. Nano Energy 2019, 65, 104027. [Google Scholar]
- (23).Romieu A; Dejouy G; Valverde IE Quest for novel fluorogenic xanthene dyes: Synthesis, spectral properties and stability of 3-imino-3H-xanthen-6-amine (pyronin) and its silicon analog. Tetrahedron Lett. 2018, 59, 4574–4581. [Google Scholar]
- (24).Gao Z; Sharma KK; Andres AE; Walls B; Boumelhem F; Woydziak ZR; Peterson BR Synthesis of a fluorinated pyronin that enables blue light to rapidly depolarize mitochondria. RSC Med. Chem 2022, 13, 456–462. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Wang C; Chi W; Qiao Q; Tan D; Xu Z; Liu X. Twisted intramolecular charge transfer (TICT) and twists beyond TICT: from mechanisms to rational designs of bright and sensitive fluorophores. Chem. Soc. Rev 2021, 50, 12656–12678. [DOI] [PubMed] [Google Scholar]
- (26).Wang C; Qiao Q; Chi W; Chen J; Liu W; Tan D; McKechnie S; Lyu D; Jiang X-F; Zhou W; Xu N; Zhang Q; Xu Z; Liu X. Quantitative Design of Bright Fluorophores and AIEgens by the Accurate Prediction of Twisted Intramolecular Charge Transfer (TICT). Angew. Chem., In. Ed 2020, 59, 10160–10172. [DOI] [PubMed] [Google Scholar]
- (27).Grimm JB; Muthusamy AK; Liang Y; Brown TA; Lemon WC; Patel R; Lu R; Macklin JJ; Keller PJ; Ji N; Lavis LD A general method to fine-tune fluorophores for live-cell and in vivo imaging. Nat. Methods 2017, 14., 987–994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Grimm JB; English BP; Chen J; Slaughter JP; Zhang Z; Revyakin A; Patel R; Macklin JJ; Normanno D; Singer RH; Lionnet T; Lavis LD A general method to improve fluorophores for live-cell and single-molecule microscopy. Nat. Methods 2015, 12, 244–250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Sun W-C; Gee KR; Klaubert DH; Haugland RP Synthesis of fluorinated fluoresceins. J. Org. Chem 1997, 62, 6469–6475. [Google Scholar]
- (30).Mitronova GY; Belov VN; Bossi ML; Wurm CA; Meyer L; Medda R; Moneron G; Bretschneider S; Eggeling C; Jakobs S; Hell SW New Fluorinated Rhodamines for Optical Microscopy and Nanoscopy. Chemistry 2010, 16, 4477–4488. [DOI] [PubMed] [Google Scholar]
- (31).Calculator Plugins were used for structure property prediction and calculation, Marvin 18.30.0, 2018, ChemAxon. [Google Scholar]
- (32).Becke AD Density-Functional Thermochemistry .3. the Role of Exact Exchange. J. Chem. Phys 1993, 98, 5648–5652. [Google Scholar]
- (33).Grimme S; Antony J; Ehrlich S; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements. H-Pu. J. Chem. Phys 2010, 132, 19. [DOI] [PubMed] [Google Scholar]
- (34).Grimme S; Ehrlich S; Goerigk L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem 2011, 32, 1456–1465. [DOI] [PubMed] [Google Scholar]
- (35).Lee CT; Yang WT; Parr RG Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron-Density. Phys. Rev. B 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
- (36).Schafer A; Huber C; Ahlrichs R. Fully Optimized Contracted Gaussian-Basis Sets of Triple Zeta Valence Quality for Atoms Li to Kr. J. Chem. Phys 1994, 100, 5829–5835. [Google Scholar]
- (37).Barone V; Cossi M. Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar]
- (38).Cossi M; Rega N; Scalmani G; Barone V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem 2003, 24, 669–681. [DOI] [PubMed] [Google Scholar]
- (39).Klamt A; Schuurmann G. Cosmo - A New Approach to Dielectric SCreening in Solvents with Explicit Expressions for the SCreening Energy and its Gradient. J. Chem. Soc.-Perkin Trans. 2 1993, 799–805. [Google Scholar]
- (40).Neese F; Wennmohs F; Becker U; Riplinger C. The ORCA quantum chemistry program package. J. Chem. Phys 2020, 152, 18. [DOI] [PubMed] [Google Scholar]
- (41).Neese F. Software update: the ORCA program system, Version 4.0. Wiley Interdiscip. Rev.-Comput. Mol. Sci 2018, 8, 6. [Google Scholar]
- (42).Neese F. Software update: The ORCA program system-Version 5.0. Wiley Interdiscip. Rev.-Comput. Mol. Sci 2022, e1606. [Google Scholar]
- (43).Hirata S; Head-Gordon M. Time-dependent density functional theory within the Tamm–Dancoff approximation. Chem. Phys. Lett 1999, 314, 291–299. [Google Scholar]
- (44).Savarese M; Raucci U; Adamo C; Netti P; Ciofini I; Rega N. Non-Radiative Decay Paths in Rhodamines: New Theoretical Insights. Phys. Chem. Chem. Phys 2014, 16, 20681–20688. [DOI] [PubMed] [Google Scholar]
- (45).Woydziak ZR; Fu L; Peterson BR, Synthesis of Fluorinated Benzophenones, Xanthones, Acridones, and Thioxanthones by Iterative Nucleophilic Aromatic Substitution. J. Org. Chem 2011, 77, 473–481. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).Arambula C; Rodrigues J; Koh JJ; Woydziak Z. Synthesis of Rhodamines and Rosamines Using 3,6-Difluoroxanthone as a Common Intermediate. J. Org. Chem 2021, 86, 17856–17865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (47).Ferrer FJA; Santoro F. Comparison of vertical and adiabatic harmonic approaches for the calculation of the vibrational structure of electronic spectra. Phys. Chem. Chem. Phys 2012, 14, 13549–13563. [DOI] [PubMed] [Google Scholar]
- (48).Sirohiwal A; Berraud-Pache R; Neese F; Izsak R; Pantazis DA Accurate Computation of the Absorption Spectrum of Chlorophyll alpha with Pair Natural Orbital Coupled Cluster Methods. J. Phys. Chem. B 2020, 124, 8761–8771. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Berraud-Pache R; Neese F; Bistoni G; Izsak R. Computational Design of Near-Infrared Fluorescent Organic Dyes Using an Accurate New Wave Function Approach. J. Phys. Chem. Lett 2019, 10, 4822–4828. [DOI] [PubMed] [Google Scholar]
- (50).de Souza B; Farias G; Neese F; Izsak R. Predicting Phosphorescence Rates of Light Organic Molecules Using Time-Dependent Density Functional Theory and the Path Integral Approach to Dynamics. J. Chem. Theory Comput. 2019, 15, 1896–1904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).de Souza B; Neese F; Izsak R. On the theoretical prediction of fluorescence rates from first principles using the path integral approach. J. Chem. Phys 2018, 148, 10. [DOI] [PubMed] [Google Scholar]
- (52).Chai J-D; Head-Gordon M. Systematic optimization of long-range corrected hybrid density functionals. J. Chem. Phys 2008, 128, 084106/084101. [DOI] [PubMed] [Google Scholar]
- (53).Grimme S; Antony J; Ehrlich S; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys 2010, 132, 154104/154101. [DOI] [PubMed] [Google Scholar]
- (54).Perdew JP; Burke K; Ernzerhof M. Generalized gradient approximation made simple. Phys. Rev. Lett 1996, 77, 3865–3868. [DOI] [PubMed] [Google Scholar]
- (55).Becke AD Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A: Gen. Phys 1988, 38, 3098. [DOI] [PubMed] [Google Scholar]
- (56).Berraud-Pache R; Neese F; Bistoni G; Izsák R. Unveiling the Photophysical Properties of Boron-dipyrromethene Dyes Using a New Accurate Excited State Coupled Cluster Method. J. Chem. Theory Comput. 2020, 16, 564–575. [DOI] [PubMed] [Google Scholar]
- (57).Chantzis A; Cerezo J; Perrier A; Santoro F; Jacquemin D. Optical Properties of Diarylethenes with TD-DFT: 0–0 Energies, Fluorescence, Stokes Shifts, and Vibronic Shapes. J. Chem. Theory Comput. 2014, 10, 3944–3957. [DOI] [PubMed] [Google Scholar]
- (58).Grimm JB; Muthusamy AK; Liang Y; Brown TA; Lemon WC; Patel R; Lu R; Macklin JJ; Keller PJ; Ji N; Lavis LD A general method to fine-tune fluorophores for live-cell and in vivo imaging. Nat. Methods 2017, 14, 987–994. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





