Next Article in Journal
Optimization of Extraction Conditions and Cytotoxic Activity of Rapanone in Comparison to Its Homologue, Embelin
Next Article in Special Issue
Donor–Acceptor Complexes of (5,10,15,20-Tetra(4-methylphenyl)porphyrinato)cobalt(II) with Fullerenes C60: Self-Assembly, Spectral, Electrochemical and Photophysical Properties
Previous Article in Journal
Effect of Hop Varieties and Forms in the Hopping Process on Non-Alcoholic Beer Quality
Previous Article in Special Issue
Azocalix[4]arene-Rhodamine Supramolecular Hypoxia-Sensitive Systems: A Search for the Best Calixarene Hosts and Rhodamine Guests
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Amine-Reactive BODIPY Dye: Spectral Properties and Application for Protein Labeling

by
Ksenia V. Ksenofontova
1,*,
Anastasia A. Kerner
1,
Alexander A. Ksenofontov
1,2,
Artyom Yu. Shagurin
3,
Pavel S. Bocharov
2,
Michael M. Lukanov
1,2,
Airat R. Kayumov
4,
Darya E. Zhuravleva
4,
Zalina I. Iskhakova
4,
Evgeniy E. Molchanov
5,
Dmitriy A. Merkushev
1,
Ilya A. Khodov
2 and
Yuriy S. Marfin
1,6
1
Department of Inorganic Chemistry, Ivanovo State University of Chemistry and Technology, 10, Sheremetevskiy pr., 153000 Ivanovo, Russia
2
G. A. Krestov Institute of Solution Chemistry of The Russian Academy of Sciences, 1, ul. Akademicheskaya, 153045 Ivanovo, Russia
3
Laboratoire de Spectroscopie pour les Interactions, la Réactivité et l’Environnement (UMR CNRS 8516), Université de Lille, CEDEX, 59655 Villeneuve d’Ascq, France
4
Institute of Fundamental Medicine and Biology, Kazan Federal University, 18, ul. Kremlyovskaya, 420008 Kazan, Russia
5
Department of Fine Organic Synthesis Technology, Ivanovo State University of Chemistry and Technology, 7, Sheremetevskiy pr., 153000 Ivanovo, Russia
6
Pacific National University, 136, ul. Tihookeanskaya, 680042 Khabarovsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(22), 7911; https://doi.org/10.3390/molecules27227911
Submission received: 25 October 2022 / Revised: 11 November 2022 / Accepted: 13 November 2022 / Published: 16 November 2022
(This article belongs to the Special Issue Molecular Recognition of Host/Guest Molecules)

Abstract

:
A boron-dipyrromethene (BODIPY) derivative reactive towards amino groups of proteins (NHS-Ph-BODIPY) was synthesized. Spectroscopic and photophysical properties of amine-reactive NHS-Ph-BODIPY and its non-reactive precursor (COOH-Ph-BODIPY) in a number of organic solvents were investigated. Both fluorescent dyes were characterized by green absorption (521–532 nm) and fluorescence (538–552 nm) and medium molar absorption coefficients (46,500–118,500 M−1·cm−1) and fluorescence quantum yields (0.32 – 0.73). Solvent polarizability and dipolarity were found to play a crucial role in solvent effects on COOH-Ph-BODIPY and NHS-Ph-BODIPY absorption and emission bands maxima. Quantum-chemical calculations were used to show why solvent polarizability and dipolarity are important as well as to understand how the nature of the substituent affects spectroscopic properties of the fluorescent dyes. NHS-Ph-BODIPY was used for fluorescent labeling of a number of proteins. Conjugation of NHS-Ph-BODIPY with bovine serum albumin (BSA) resulted in bathochromic shifts of absorption and emission bands and noticeable fluorescence quenching (about 1.5 times). It was demonstrated that the sensitivity of BSA detection with NHS-Ph-BODIPY was up to eight times higher than with Coomassie brilliant blue while the sensitivity of PII-like protein PotN (PotN) detection with NHS-Ph-BODIPY and Coomassie brilliant blue was almost the same. On the basis of the molecular docking results, the most probable binding sites of NHS-Ph-BODIPY in BSA and PotN and the corresponding binding free energies were estimated.

Graphical Abstract

1. Introduction

Detection, tracking, and imaging of proteins are considerable challenges for modern science. One of the ways of tackling this challenge is to label proteins with fluorescent dyes. Fluorescent labels allow imaging various biochemical processes including protein interactions with other biomolecules, protein localization, protein dynamics, enzyme activity, etc. [1,2].
Currently, three main classes of fluorescent labels are widely used [3]: organic dyes, fluorescent proteins, and quantum dots. Each of them has its own advantages and disadvantages. In this regard, the development of new fluorophores that meet all the requirements of bioimaging continues unabated.
Organic dyes being the most popular fluorescent labels include fluoresceins [4], rhodamines [5,6], coumarins [7], dipyrrin complexes [3,8,9,10], pyrenes [11], cyanines [12], etc. The most important feature of organic dyes is the possibility of tuning their fluorescent properties by means of chemical modification and/or environmental change.
Bright members of organic dyes are boron-dipyrromethene (BODIPY) fluorophores and their derivatives [13]. This family of dyes possesses a number of striking properties, such as great stability, high molar absorption coefficients, excellent fluorescence quantum yields, relatively small Stokes shifts, and sharp excitation and emission peaks [14]. Moreover, its ease of chemical modification opens up great possibilities for a synthesis of new fluorescent compounds with predetermined properties.
A rather new promising area in BODIPYs chemistry is the development of dyes reactive towards principal functional groups of proteins [15,16,17,18,19,20,21,22,23,24]. Such fluorescent dyes possess certain reactive substituents capable of covalent binding with proteins, thereby providing their effective labeling. Unfortunately, most of existing reactive BODIPY derivatives have several issues with solubility in water and/or photophysical characteristics. Therefore, the search for reactive fluorescent dyes suitable for conjugation with proteins continues.
In this work, we synthesized a BODIPY derivative reactive towards amino groups of proteins. A comparative analysis of spectroscopic and photophysical properties as well as solvatochromic behavior of the synthesized reactive fluorescent dye and its non-reactive precursor was carried out by means of ultraviolet-visible spectroscopy and steady-state and time-resolved fluorescence spectroscopy supported by quantum chemical calculations. The investigated reactive fluorescent dye was used for labeling of a range of proteins. The sum of the experimental and theoretical results suggests that the synthesized amine-reactive fluorescent dye is a promising fluorescent label for molecular biology and biotechnology.

2. Results and Discussion

2.1. Synthesis of Reactive BODIPY

The amine-reactive fluorescent dye succinimidyl ester of 4-(2,6-diethyl-4,4-difluoro-1,3,5,7-tetramethyl-4-bora-3a,4a-diaza-s-indacene-8-yl)-benzoic acid (NHS-Ph-BODIPY) was synthesized from 4-(2,6-diethyl-4,4-difluoro-1,3,5,7-tetramethyl-4-bora-3a,4a-diaza-s-indacene-8-yl)-benzoic acid (COOH-Ph-BODIPY) following the procedure [25] (Scheme 1). Synthetic details and characterization of the compounds obtained are given in Section 3.3.

2.2. UV-Vis and Fluorescence Spectroscopy Experiments of BODIPYs

Absorption spectra, emission spectra, fluorescence decay curves as well as a number of photophysical characteristics of synthesized fluorescent dyes COOH-Ph-BODIPY and NHS-Ph-BODIPY in a range of organic solvents of various nature were obtained (Figure 1 and Figure S1, Table 1).
In Table 1, λabs max is the maximum absorption wavelength, λem max is the maximum emission wavelength, Δν is the Stokes shift, and ε is the molar absorption coefficient at the maximum absorption wavelength. The asterisk (*) indicates the average fluorescence lifetime τ ¯ calculated using Equation (1) [26]:
τ ¯ = α 1 τ 1 2 + α 2 τ 2 2 α 1 τ 1 + α 2 τ 2
where α is the amplitude.
The absorption and emission spectra of COOH-Ph-BODIPY and NHS-Ph-BODIPY are typical for this class of dyes (Figure 1). The absorption spectra gave two bands in the ranges of 521–532 and 375–385 nm due to S0–S1 and S0–S2 electron transitions, respectively, while the emission spectra gave one band in the region of 538–552 nm due to S1–S0 electron transitions (Section 2.3). The absorption and emission bands maxima of NHS-Ph-BODIPY are 2–7 nm red shifted compared with the spectra of COOH-Ph-BODIPY. The fluorescent dyes are characterized by medium molar absorption coefficients and fluorescence quantum yields which change non-linearly from COOH-Ph-BODIPY to NHS-Ph-BODIPY. The fluorescence lifetimes of NHS-Ph-BODIPY are shorter than those of COOH-Ph-BODIPY, while the radiative and non-radiative rate constants change non-linearly.
Solvent effects on spectral characteristics of COOH-Ph-BODIPY and NHS-Ph-BODIPY, namely the maximum absorption νabs(max) and emission νem(max) wavenumbers and Stokes shift Δν, were analyzed in terms of the Kamlet–Taft [27], Catalán [28], and Lippert–Mataga [29] equations.
Table 2 and Table S1 list the estimated regression coefficients y0, ad and their standard errors as well as coefficients of determination R2 for the multiple linear regression analyses of the maximum absorption νabs(max) and emission νem(max) wavenumbers and Stokes shifts Δν of COOH-Ph-BODIPY and NHS-Ph-BODIPY according to Equations (5) and (S2) for the solvents chosen.
The analyses of the νabs(max) and νem(max) data within the Kamlet–Taft model, in which solvent polarizability and dipolarity effects are combined in the single parameter π*, show poor fits with the R2 = 0.44–0.66 and large standard errors on the estimated regression coefficients aα, bβ, and cπ* as goodness-of-fit criteria (Table S1). In contrast, the analyses of the same data within the Catalán model, in which solvent polarizability and dipolarity effects are separated by the two parameters SP and SdP, give good to perfect fits with R2 = 0.75–0.96 (Table 2). To visualize the goodness-of-fits of νabs(max) and νem(max) as functions of the Catalán solvent parameters {SA, SB, SP, SdP}, the plots of νabs(max) and νem(max) of COOH-Ph-BODIPY and NHS-Ph-BODIPY calculated according to Equation (5) using the estimated values of y0, aSA, bSB, cSP, and dSdP vs. the corresponding experimental νabs(max) and νem(max) values were drawn (Figure 2).
The Catalán model further reveals solvent properties that are mainly responsible for the observed spectral shifts. The very large (negative) estimated regression coefficients cSP compared to the aSA, bSB, and dSdP ones as well as the relatively large standard errors on the aSA, bSB, and dSdP compared to those on the cSP (Table 2) indicate that the changes of νabs(max) and νem(max) may primarily reflect changes in polarizability of the environment of COOH-Ph-BODIPY and NHS-Ph-BODIPY [30]. The other supporting evidence comes from the multiple linear regression analyses of the νabs(max) and νem(max) according to Equation (5) with {SB, SP, SdP}, {SA, SP, SdP}, and {SA, SB, SP} as independent variables (Table S2). The analyses give good fits with an R2 equal to 0.91, 0.89, and 0.72 and 0.75, 0.69, and 0.67 for νabs(max) and νem(max) of COOH-Ph-BODIPY, respectively, as well as 0.91, 0.92, and 0.78 and 0.85, 0.82, and 0.90 for νabs(max) and νem(max) of NHS-Ph-BODIPY, respectively. It is significant to note that the lowest coefficients of determination (indicated in bold) were obtained in case of {SA, SB, SP} as independent variables (except for νem(max) of NHS-Ph-BODIPY), indicating that solvent dipolarity should not be neglected as an influencing factor. In contrast, the multiple linear regression analyses of the νabs(max) and νem(max), according to Equation (5) with {SA, SB, SdP} as independent variables (Table S2), show poor fits with an R2 equal to 0.54 and 0.52 for νabs(max) and νem(max) of COOH-Ph-BODIPY, respectively, as well as 0.38 and 0.39 for νabs(max) and νem(max) of NHS-Ph-BODIPY, respectively. Thus, the crucial solvent property affecting the maximum absorption νabs(max) and emission νem(max) wavenumbers of COOH-Ph-BODIPY and NHS-Ph-BODIPY is solvent polarizability with a small contribution of solvent dipolarity.
The Lippert plots of the Stokes shifts Δν of COOH-Ph-BODIPY and NHS-Ph-BODIPY vs. orientation polarizabilities Δf of the solvents chosen are presented in Figure S2. There are extremely poor linear relationships between Δν and Δf with an R2 equal to 0.03 and 0.05, respectively. The small slopes (−68 and 56 cm−1, respectively) imply that the dipole moments of COOH-Ph-BODIPY and NHS-Ph-BODIPY do not change noticeably between the ground and excited states [31]. The other supporting evidence comes from the multiple linear regression analyses of the Δν within the Kamlet–Taft and Catalán models (Table S1 and Table 2), both of which show poor fits with the R2 equal to 0.31 and 0.29 for COOH-Ph-BODIPY, respectively, as well as 0.61 and 0.70 for NHS-Ph-BODIPY, respectively. It is interesting to note that quantum chemical calculations gave another explanation of the abovementioned trends (Section 2.3).

2.3. Quantum Chemical Calculations of BODIPYs

In order to gain more insight into the abovementioned relations, a series of quantum chemical calculations was performed. First, a semiempirical metadynamic screening revealed the existence of four different conformers for both COOH-Ph-BODIPY and NHS-Ph-BODIPY, of which only two were found to be non-redundant according to further DFT optimization. Those conformations differ only in the relative orientation of ethyl groups, which have limited impact on spectroscopic properties. As was later confirmed, all conformers have approximately the same excitation energies (Figure S3). It is, thus, reasonable to expect that existence of those conformers does not lead to any complications and further analysis can be performed only on the most thermodynamically stable conformer (Figure 3).
Firstly, we have obtained vibrationally resolved absorption and emission spectra of the studied molecules. Calculations were performed using the VG-FC model. As shown in the literature [32], neglect of the excited state geometry relaxation is not only a reasonable, but sometimes crucially important (for calculation convergence) assumption. The comparison of the ground and excited state geometries (Figure 3) shows that although there is a noticeable change in structure between S0 and S1 (reduction of C7-C8-C1Ph-C2Ph dihedral angle leading to improved conjugation between BODIPY and C8-substituent), this change is consistent between the two studied dyes, so the general trends for relevant properties should be preserved.
As can be seen from the VG-FC absorption and emission spectra (Figure 4), one of the preserved trends is the presence of a reasonably slight bathochromic shift resulting from the addition of succinimidyl group to BODIPY core. The presented theoretical spectra also give the same overall shape and Stokes shifts seen in experimental data (Figure 1).
Next, we have compared the results of the multiple linear regression analysis (Section 2.2) with the data obtained from the quantum chemical calculations (Table 3). In particular, relatively large (~3.6 D) ground state dipole moments of both COOH-Ph-BODIPY and NHS-Ph-BODIPY as well as large (~1.5 D) dipole moment changes upon excitation show why such Catalán solvent parameters as polarizability SP and dipolarity SdP dominate the abovementioned regressions. Moreover, the poor applicability of the Lippert relations may be explained not by a small dipole moment change upon excitation but by an indifference of the change towards solvent nature. Indeed, for both COOH-Ph-BODIPY and NHS-Ph-BODIPY, the dipole moment change is about 0.71 a.u. and for both non-polar n-hexane and slightly polar DMSO, the dipole moment magnitude change is about 0.68 a.u.
To gain a better understanding of the nature of this excitation, we have opted to compute the partitioned transition density matrix [33]. This technique has been used to great effect to explain the similarities and differences between compounds [34,35]. Moreover, recent papers [36] have even compared different DFT functionals in terms of the quality of their transition density matrices.
We have defined five common units for COOH-Ph-BODIPY and NHS-Ph-BODIPY—two pyrrole rings (L-pyr and R-pyr), coordination center (CC), C8-carbon atom (C8), and benzoyl group (C8-sub)—as well as the additional one for NHS-Ph-BODIPY—succinimidyl group (NHS) (Figure 5a).
As can be seen from the visual representation of the matrices (Figure 5b,c), the first singlet excitation is localized mostly on the pyrrole rings and has a local character, although there is also a significant electron transfer between the rings, which, nevertheless, is symmetric. The change in the dipole moment arises due to the density transfer between the pyrrole rings as well as C8-carbon atom and its substituent. However, the position of the phenyl ring relative to dipyrrin is almost perpendicular due to the existence of the methyl groups in 1- and 7-positions of BODIPY core. As such, steric factors around this position should prevent any solvent molecules from getting close enough to provide a significant local impact on electron density via polarization.
Comparing two matrices, it becomes clear that the nature of the phenyl substituent does not play a strong role when it comes to the first excited state–transition density on the succinimidyl group, as well as density change between it and other units, are small. As such, as will be shown below (Section 2.4 and Section 2.5), any alterations in the absorption and emission bands positions of COOH-Ph-BODIPY and NHS-Ph-BODIPY upon conjugation with proteins will come as a result of intermolecular interactions, such as polarization due to nearby amino acids or loss of polarization induced by solvent.

2.4. Fluorescent Labeling of Proteins

The synthesized amine-reactive fluorescent dye NHS-Ph-BODIPY was used for labeling of a range of proteins, such as bovine serum albumin (BSA), PII-like protein PotN (PotN), PotA subunit of polyamine ABC transporter (PotAc), and glutamine synthetase (GS).
The fluorescent labeling is based on a conjugation reaction between NHS-Ph-BODIPY possess an active succinimidyl ester moiety and a protein possess amine-containing residues (i.e., ε-amines of lysine side chains and α-amines at N-termini) to form a stable amide linkage (Scheme 2). The conjugation reaction conditions, namely slightly basic pH and low temperature, were chosen to ensure amine groups in proteins were unprotonated and to prevent hydrolysis of NHS-Ph-BODIPY to remove a reactive group [37]. Labeling details are given in Section 3.4.
At the first stage, the ability of NHS-Ph-BODIPY to label biomolecules was tested using a globular protein BSA as a model compound.
Absorption spectra, emission spectra, fluorescence decay curves, as well as a number of photophysical characteristics of NHS-Ph-BODIPY and its conjugate with BSA (BSA–NHS-Ph-BODIPY) were obtained in a mixture of DMSO and bicarbonate buffer with pH 8.3 (1:9) (Figure 6 and Figure S4, Table 4).
The absorption and emission spectra of BSA–NHS-Ph-BODIPY have several distinctive features (Figure 6). The absorption spectrum gave three bands with the maxima at 280, 380, and 530 nm, while the emission spectrum gave two bands with the maxima at 350 (not shown) and 554 nm. The new bands in the UV region are certain to relate with the absorption and emission of the protein moiety of the conjugate. The characteristic absorption and emission bands maxima of the BODIPY moiety of the conjugate are red shifted compared with the spectra of the free dye. Along with this, it is observed a decrease (about 1.5 times) of the fluorescence quantum yield of BSA–NHS-Ph-BODIPY in contrast to that of NHS-Ph-BODIPY as well as a change of the intensity decay from single exponential in case of NHS-Ph-BODIPY to double exponential in case of BSA–NHS-Ph-BODIPY. These phenomena can be explained by a dynamic quenching by aromatic amino acids located at the immediate environment of the dye in the conjugate [38,39], which is confirmed via synchronous fluorescence spectroscopy (Figure S5) and molecular docking (Section 2.5). As regards the synchronous fluorescence spectroscopy, it was found that the change from a pure BSA to a labeled BSA results in a decrease of fluorescence intensities (from 3 times in case of Δλ = 15 nm to 10 times in case of Δλ = 60 nm) and a blue shift (from 2 nm in case of Δλ = 15 nm to 10 nm in case of Δλ = 60 nm) of the bands maxima at both wavelength shifts. This suggests that the addition of NHS-Ph-BODIPY causes changes in the environment of tyrosine and tryptophan residues of BSA, namely an increase of its hydrophobicity [40].
Furthermore, on the basis of the absorption spectra of NHS-Ph-BODIPY and BSA–NHS-Ph-BODIPY, the degree of labeling of BSA was estimated (Section 3.7). The value amounts to 1 which means that each protein molecule possesses one fluorescent label.
At the next stage, the possibility of NHS-Ph-BODIPY utilization in protein analysis was examined.
To evaluate a sensitivity of protein detection with NHS-Ph-BODIPY, the double dilutions (from 10.0 to 0.1 µg) of fluorescently labeled BSA and PotN were prepared and separated with SDS-PAGE. As can be seen from Figure 7, NHS-Ph-BODIPY-labeled BSA is detected at the concentrations up to eight times lower than Coomassie-stained BSA, while both NHS-Ph-BODIPY-labeled and Coomassie-stained PotNs are found at about the same concentration range. This phenomenon can be explained by a higher probability of NHS-Ph-BODIPY to bind lysine residues at the BSA binding site than in the PotN one (Section 2.5). Next, to evaluate whether NHS-Ph-BODIPY changes the charge of a protein molecule, four proteins (PotN, PotAc, GS, BSA) with various isoelectric points (pI) and molecular weights (MW) were separated under non-denaturing conditions. No difference in electrophoretic mobility of the chosen proteins was observed (Figure 8), suggesting no significant effect of NHS-Ph-BODIPY on the physico-chemical properties of proteins.

2.5. Molecular Docking of Protein–BODIPY Conjugates

To gain insight into protein–BODIPY interactions, a two-stage molecular docking of BSA and PotN with NHS-Ph-BODIPY was carried out.
At the first stage, a blind docking was performed to determine the most probable binding site of NHS-Ph-BODIPY in BSA. It was shown that the fluorescent dye is localized in a cavity between the IB, IIA, and IIB subdomains of the protein (Figure 9a) with a binding free energy of −26.1 kJ/mol. It is noteworthy that these subdomains are known to contain major drug binding sites [41,42,43]. The amino acid composition of the binding site of NHS-Ph-BODIPY in BSA is shown in Figure 9b. The binding site was found to contain three lysine residues (Lys187, Lys221, Lys294), each of which was selected for a subsequent covalent docking. In addition, it contains tyrosine (Tyr149, Tyr156, Tyr451) and tryptophan (Trp213) residues, which is consistent with the results of synchronous fluorescence spectroscopy (Section 2.4).
At the second stage, a covalent docking was performed to estimate a binding free energy of NHS-Ph-BODIPY and BSA bound together via an amide linkage through one of the lysine residues (Lys187, Lys221, Lys294) found in the binding site at the previous stage (Figure 10 and Figure S6). The binding free energies values of Lys187-BSA–NHS-Ph-BODIPY, Lys221-BSA–NHS-Ph-BODIPY, and Lys294-BSA–NHS-Ph-BODIPY systems being −127, −129, and −108 kJ/mol, respectively, indicate an exergonic type of the conjugation reaction. The first two systems are characterized by almost equal binding free energies, which does not allow one to get an unambiguous answer about an anchor residue involved in the conjugation. Therefore, it was estimated a distance between the tyrosine and tryptophan residues (Tyr149, Tyr156, Tyr451, Trp213) and the lysine residues bound to NHS-Ph-BODIPY. The distance values are 5.669 (average) and 12.201 Å for Tyr Lys187-BSA–NHS-Ph-BODIPY and Trp···Lys187-BSA–NHS-Ph-BODIPY, respectively, and 3.524 (average) and 3.357 Å for Tyr···Lys221-BSA–NHS-Ph-BODIPY and Trp···Lys221-BSA–NHS-Ph-BODIPY, respectively. Thus, a synergy of synchronous fluorescence spectroscopy and molecular docking made it possible to establish that it is Lys221 that participate in the conjugation reaction between NHS-Ph-BODIPY and BSA.
A similar two-stage approach was used for molecular docking of PotN with NHS-Ph-BODIPY. A blind docking results shown that the fluorescent dye is localized near the B-loop of the protein (Figure 11) with a binding free energy of −31.9 kJ/mol. The binding site was found to contain one lysine residue (Lys92). A covalent docking results (Figure S7) gave a binding free energy value of −122 kJ/mol for the Lys92-PotN–NHS-Ph-BODIPY system.
It is noteworthy that the molecular docking results give explanation for a difference in sensitivity of BSA and PotN detection with NHS-Ph-BODIPY (Section 2.4). As stated above, the BSA binding site contains three lysine residues, while the PotN binding site contains only one lysine residue. Thus, lower detection limit for BSA than for PotN may be due to a higher probability of NHS-Ph-BODIPY to bind lysine residues at the binding site of the first protein than the second one.

3. Materials and Methods

3.1. Materials

Reagents for BODIPYs synthesis were purchased from Sigma-Aldrich (Moscow, Russia), with the exception of boron trifluoride diethyl etherate (BF3·OEt2) purchased from Merck Millipore (Burlington, MA, USA). Bovine serum albumin (BSA) was purchased from Sigma-Aldrich (Moscow, Russia), while StrepII-tagged PII-like protein PotN (PotN) and His6-tagged C-terminal domain of PotA subunit of polyamine ABC transporter (PotAc) from Lentilactobacillus hilgardii as well as StrepII-tagged glutamine synthetase (GS) from Bacillus subtilis were obtained as described earlier [44,45]. Organic solvents were bought from Sigma-Aldrich (Moscow, Russia), EKOS-1, and Khimmed. Substances for buffers preparation were bought from Sigma-Aldrich and Reakhim (Moscow, Russia).
Dialysis tubing MD25 with 8.0–14.0 kDa molecular weight cut-off was obtained from a commercial source and prepared for use according to the procedure [46].

3.2. Instruments

1H and 11B nuclear magnetic resonance (NMR) spectra were recorded on a Bruker Avance III 500 NMR spectrometer (Billerica, MA, USA) with operating frequencies of 500.17 for 1H and 160.48 MHz for 11B. Deuterated chloroform was chosen as a solvent for samples preparation. Tetramethylsilane was used as an internal reference (δ = 0.00 ppm) for 1H NMR studies, while BF3·OEt2 was used as an external reference (δ = 0.00 ppm) for 11B NMR studies. The following abbreviations are used to designate peak multiplicities and descriptors: s for singlet, d for doublet, t for triplet, and q for quadruplet. Infrared (IR) spectra were obtained by means of a Shimadzu IRAffinity-1 Fourier transform IR spectrophotometer (Billerica, MA, USA) equipped with a Specac Quest ATR Diamond GS10800-B accessory in the mid-infrared (400–4000 cm−1) region. The following abbreviations are used to designate signal intensities: w for weak, m for medium, s for strong, br for broad, and sh for sharp. Matrix-assisted laser desorption/ionization (MALDI) time of flight (TOF) mass spectra (MS) were recorded on a Shimadzu AXIMA Confidence MALDI TOF-TOF mass spectrometer in positive ion reflectron mode.
Ultraviolet-visible (UV-Vis) spectroscopy experiments were performed on an Aquilon SF-104 spectrophotometer (Waltham, MA, USA). Absorption spectra were measured in the range of 190–800 nm. Fluorescence spectroscopy experiments were performed on an Agilent Cary Eclipse fluorescence spectrophotometer (Santa Clara, CA, USA). Steady-state emission spectra were measured in the ranges of 270–400 and 490–800 nm with excitation wavelengths of 260 and 480 nm, respectively. Synchronous emission spectra were measured in the range of 250–350 nm with wavelength shifts Δλ of 15 and 60 nm. Excitation and emission slit widths values were 5.0 nm. Time-resolved fluorescence spectroscopy experiments were carried out by means of a PicoQuant FluoTime 300 high performance fluorescence lifetime and steady state spectrometer with a PicoQuant PLS 450 sub-nanosecond pulsed light-emitting diode as an excitation source. An instrument response function of a system was measured with a stray light signal of a dilute colloidal silica suspension (LUDOX®). Fluorescence decay curves were measured at the maximum of the emission peaks and fluorescence lifetimes were obtained by reconvolution of the decay curves using a PicoQuant EasyTau 2 software package. Standard quartz cuvettes with 10 mm light path were used for all experiments.
Labeled proteins were separated with sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS-PAGE) and basic native–polyacrylamide gel electrophoresis (basic native-PAGE) [47], visualized with Bio-Rad ChemiDoc XRS+ System in trans-UV mode and then stained with Coomassie brilliant blue.

3.3. Synthesis of Reactive BODIPY

4-(2,6-Diethyl-4,4-difluoro-1,3,5,7-tetramethyl-4-bora-3a,4a-diaza-s-indacene-8-yl)-benzoic acid (COOH-Ph-BODIPY). Synthesis was carried out using a mechanochemical approach [48] (Scheme 1, Figures S8–S11). 3-Ethyl-2,4-dimethylpyrrole (0.3 mL, 2.2 mmol, 2 equiv.), 4-carboxybenzaldehyde (167.0 mg, 1.1 mmol, 1 equiv.), trifluoroacetic acid (TFA) (1 drop), p-chloranil (407.0 mg, 1.7 mmol, 1.5 equiv.), and a small amount of dichloromethane (DCM) as a binder additive were mixed with a pestle and mortar. The mixture was ground for 10 min until a thick dark purple paste with green metallic luster was formed. After that, triethylamine (Et3N) (1.2 mL, 9.0 mmol, 8 equiv.) and dropwise BF3·OEt2 (1.3 mL, 10.3 mmol, 9 equiv.) were added. The mixture was ground for 10 min until a thick dark purple paste with green metallic luster was formed. After that, the resulting paste was dissolved in DCM–ethyl acetate (1:1) mixture (150 mL), filtered through a Buchner funnel, and washed with brine (3 × 50 mL). The organic solvents were removed in vacuo, and the crude residue was purified by column chromatography on silica gel using diethyl ester–hexane (1:1) mixture. After the removal of eluent in vacuo, 131.8 mg of dark red powder of the desired product was obtained, resulting in 28% yield. 1H NMR: δ (ppm) 8.23 (d, 2H, J = 7.9 Hz, –Caryl–H), 7.44 (d, 2H, J = 7.8 Hz, –Caryl–H), 2.53 (s, 6H, –CH3), 2.30 (q, 4H, J = 7.6 Hz, –CH2–), 1.26 (s, 6H, –CH3), 0.98 (t, 6H, J = 7.6 Hz, –CH3). 11B NMR: δ (ppm) 0.78 (t, 1B, J = 33.9 Hz, –BF2–). IR: ν (cm−1) 2963–2870 (m, br, Caliphatic–H, O–H), 1684 (s, sh, C=O), 1473 (m, N–B), 1261 (s, sh, C–O), 1112 (m, B–F), 1073 (m, B–F). MS (MALDI TOF): m/z calculated for C24H27BF2N2O2+ [M]+ 424.21, found to be 424.99.
Succinimidyl ester of 4-(2,6-diethyl-4,4-difluoro-1,3,5,7-tetramethyl-4-bora-3a,4a-diaza-s-indacene-8-yl)-benzoic acid (NHS-Ph-BODIPY). Synthesis was carried out on the basis of the previously described procedure [25] (Scheme 1, Figures S8–S11). After the synthesis and all workups, 15.3 mg of orange-red powder of the desired product was obtained resulting in 41% yield. 1H NMR: δ (ppm) 8.26 (d, 2H, J = 7.9 Hz, –Caryl–H), 7.49 (d, 2H, J = 7.9 Hz, –Caryl–H), 2.96 (d, 4H, J = 7.8 Hz, –CH2succinimide), 2.53 (s, 6H, –CH3), 2.30 (q, 4H, J = 7.6 Hz, –CH2–), 1.27 (s, 6H, –CH3), 0.98 (t, 6H, J = 7.5 Hz, –CH3). 11B NMR: δ (ppm) 0.76 (t, 1B, J = 34.0 Hz, –BF2–). IR: ν (cm−1) 2964–2853 (m, br, Caliphatic–H), 1800 (w, C=Osuccinimide), 1767 (m, C=Osuccinimide), 1739 (s, C=O), 1473 (m, N–B), 1117 (m, B–F), 1064 (m, B–F). MS (MALDI TOF): m/z calculated for C28H30BF2N3O4+ [M]+ 521.23, found to be 522.06.

3.4. Fluorescent Labeling of Proteins

Fluorescent labeling of proteins was carried out on the basis of the standard amine-reactive probe labeling protocol [49] (Scheme 2). A solution of NHS-Ph-BODIPY (8 equiv.) in dimethyl sulfoxide (DMSO) was slowly added to a solution of protein (1 equiv.) in bicarbonate or phosphate buffer with pH 8.3. After a thorough stirring, the mixture was incubated at 4 °C for 24 h. After completion of the reaction, the mixture was dialyzed against 70% ethanol (three times) and then distilled water (three times) to give a pure labeled protein solution.

3.5. Determination of Photophysical Characteristics

The fluorescence quantum yields Φ of the compounds investigated were estimated by comparison with Rhodamine 6G as a standard of known fluorescence quantum yield (Φ = 0.91 in ethanol [50]) using Equation (2) [50]:
Φ x = Φ s t S x S s t 1 10 A s t 1 10 A x n x 2 n s t 2
where S is the integrated area under the emission spectrum, A is the absorbance at the excitation wavelength, and n is the refractive index of the solvent. The subscripts x and st refer to the unknown and reference solutions, respectively.
The radiative kr and non-radiative knr rate constants were calculated using Equations (3) and (4) [26], respectively:
k r = Φ τ  
k n r = 1 τ k r
where τ is the fluorescence lifetime.

3.6. Description of Solvent Effects via Multiparameter Approach

In order to describe solvent effects on spectral characteristics of the compounds investigated, a multiparameter approach using the Catalán equation (Equation (5)) [28] was applied:
y = y 0 + a S A S A + b S B S B + c S P S P + d S d P S d P
where y is the value of a solvent-dependent physicochemical property in a given solvent, y0 is the statistical quantity corresponding to the value of a solvent-dependent physicochemical property in the gas phase or in an inert solvent, SA is the empirical parameter of solvent hydrogen-bond donor acidity, SB is the empirical parameter of solvent hydrogen-bond acceptor basicity, SP is the empirical parameter of solvent polarizability, SdP is the empirical parameter of solvent dipolarity, and ad are the regression coefficients describing the sensitivity of a physicochemical property y to the different solute/solvent interaction mechanisms.

3.7. Determination of Degree of Labeling

The degrees of labeling of the proteins investigated were estimated using Equation (6) [49]:
D O L = A max c o n j ε max d y e ε max p r o t A 280   c o n j A max c o n j A 280   d y e A m a x   d y e
where DOL is the degree of labeling, Amax conj and Amax dye are the absorbances of the conjugated dye and the free dye at the maximum absorption wavelength, respectively, εmax prot and εmax dye are the molar absorption coefficients of the free protein and the free dye at the absorption wavelength, respectively, and A280 conj and A280 dye are the absorbances of the labeled protein and maximum the free dye at 280 nm, respectively.

3.8. Quantum Chemical Calculations Procedure

The conformational search for COOH-Ph-BODIPY and NHS-Ph-BODIPY was performed in a metadynamic approximation. The calculations were carried out by means of the xtb program [51] with the CREST add-on [52]. The more accurate GFN2-xTB method was chosen along with the iMTD-GC algorithm. The conformational screening revealed four conformers for COOH-Ph-BODIPY and four conformers for NHS-Ph-BODIPY. The lowest energy conformers were chosen for further DFT and TDDFT study. Each conformer was optimized at the CAM-B3LYP/6-31G(d,p) level, which provides a balanced description of both ground and excited states [53]. For each resulting structure, ten excitation vertical absorption spectra were computed. For the first three excitations, gradients as well as vertical gradient Franck-Condon (VG-FC) vibronic spectra within the TD formalism as implemented in the FCclasses 3.0 program [54] were also obtained. Vibronic spectra calculations used ground state vibrational frequencies, vertical excitation energies computed at ground state geometries and excited state gradients computed at the same geometries. All empirical parameters pertaining to the VG-FC algorithm, such as number of points for the fast Fourier transform (FFT), were directly estimated by the FCclasses program. All vibronic spectra are broadened with Gaussian functions with the constant width of 322 cm−1. This value was chosen because it gives the best agreement between experimental and theoretical spectra of fully unsubstituted BODIPY. All calculations, unless otherwise mentioned, were performed within the CPCM model with DMSO parameters using the ORCA 5.0 program suite [55,56,57]. Quantities of interest, such as ground and excited state dipole moments as well as transition density matrix elements, were computed using the latest version of the Multiwfn program [58]. A visualization of the quantum chemical calculations results were made by means of both the Chemcraft 1.8 [59] and the VMD [60] programs.

3.9. Molecular Docking Procedure

The molecular docking of the proteins–BODIPY conjugates was performed in two stages. At the first stage, a blind docking was carried out by means of the Autodock 4.2 program [61]. The crystal structures of the proteins under study were taken from the Protein Data Bank: BSA [62] and PotN [63]. The structure of NHS-Ph-BODIPY was obtained by virtue of geometry optimization (Section 3.8). The calculations were performed for a 126 Å × 126 Å × 126 Å grid with BSA in its center with a step of 0.7 Å and a 90 Å × 126 Å × 126 Å grid with PotN in its center with a step of 0.4 Å. Each docking experiment included 50 runs with a maximum of 25 million energy evaluations using the Lamarckian genetic algorithm [64]. Conformations of the protein–BODIPY conjugates with minimum energies were assumed as the most stable ones. At the second stage, a covalent docking was carried out by means of the CovDock program [65]. On the basis of the blind docking results, a covalent binding of NHS-Ph-BODIPY and lysine residues (imine condensation pre-defined reaction) at the most energetically favorable binding site of the proteins was calculated. Conformations of the protein–BODIPY conjugates with minimum energies were again assumed as the most stable ones. A visualization of the molecular docking results were made by means of the UCSF Chimera program [66].

4. Conclusions

The development of fluorescent dyes reactive towards principal functional groups of protein is a promising area of modern chemistry. Therefore, we carried out a thorough investigation of spectroscopic and photophysical properties as well as solvatochromic behavior of the synthesized amine-reactive fluorescent dye NHS-Ph-BODIPY and its non-reactive precursor COOH-Ph-BODIPY by means of ultraviolet-visible and fluorescence spectroscopy supported by quantum chemical calculations that allowed us to better understand some specific aspects of their functioning in solutions. An inherent stable bright green fluorescence of NHS-Ph-BODIPY excited us to use this dye for fluorescent labeling of a number of proteins of various nature. NHS-Ph-BODIPY was demonstrated to act at least at the same level or even several times better than Coomassie brilliant blue when detecting proteins via SDS-PAGE. Thus, the sum of the experimental and theoretical results suggests that the synthesized amine-reactive fluorescent dye NHS-Ph-BODIPY is a promising fluorescent label for fluorescent immunoassay, cellular imaging and in vivo imaging, flow cytometry, immunohistochemical staining, etc.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27227911/s1, Figure S1: Fluorescence decay curves of COOH-Ph-BODIPY and NHS-Ph-BODIPY in DMSO and n-propanol; Figure S2: Lippert plots for COOH-Ph-BODIPY and NHS-Ph-BODIPY; Figure S3: First 10 singlet excitation energies and their oscillator strengths for each of two conformers of COOH-Ph-BODIPY and NHS-Ph-BODIPY according to TD-DFT calculations; Figure S4: Fluorescence decay curves of NHS-Ph-BODIPY and BSA–NHS-Ph-BODIPY in DMSO–bicarbonate buffer (pH 8.3) mixture (1:9); Figure S5: Synchronous emission spectra of BSA and BSA–NHS-Ph-BODIPY in DMSO–bicarbonate buffer (pH 8.3) mixture (1:9); Figures S6 and S7: Amino acid environment of NHS-Ph-BODIPY in BSA and PotN according to covalent docking results of Lys187-BSA–NHS-Ph-BODIPY, Lys221-BSA–NHS-Ph-BODIPY, Lys294-BSA–NHS-Ph-BODIPY, and Lys92-PotN–NHS-Ph-BODIPY systems; Figures S8–S11: 1H NMR, 11B NMR, IR, and MS spectra of COOH-Ph-BODIPY and NHS-Ph-BODIPY; Tables S1 and S2: Regression coefficients y0, ad, and coefficients of determination R2 for multiple linear regression analysis of maximum absorption νabs(max) and emission νem(max) wavenumbers and Stokes shifts Δν of COOH-Ph-BODIPY and NHS-Ph-BODIPY as a function of Kamlet–Taft {α, β, π*} and Catalán {SB, SP, SdP}, {SA, SP, SdP}, {SA, SB, SdP}, and {SA, SB, SP} solvent scale parameters; S1: Description of solvent effects.

Author Contributions

Conceptualization, K.V.K.; Methodology, K.V.K., A.A.K. (Alexander A. Ksenofontov), A.R.K. and Y.S.M.; Validation, K.V.K.; Formal Analysis, K.V.K., A.A.K. (Anastasia A. Kerner), A.A.K. (Alexander A. Ksenofontov), A.Y.S., P.S.B. and M.M.L.; Investigation, K.V.K., A.A.K. (Anastasia A. Kerner), D.E.Z., Z.I.I., E.E.M., D.A.M. and I.A.K.; Writing—Original Draft, K.V.K., A.A.K. (Alexander A. Ksenofontov) and A.Y.S.; Writing—Review and Editing, K.V.K. and Y.S.M.; Visualization, K.V.K., A.A.K. (Anastasia A. Kerner), A.A.K. (Alexander A. Ksenofontov), A.Y.S., P.S.B., D.E.Z. and Z.I.I.; Supervision, K.V.K. and Y.S.M.; Funding Acquisition, K.V.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation grant number 22-73-00337 (synthesis, UV-Vis and fluorescence spectroscopy experiments, quantum chemical calculations) and the Council for Grants of the President of the Russian Federation scholarship number SP-2910.2021.4 (fluorescent labeling, molecular docking), and was supported by the Kazan Federal University Strategic Academic Leadership Program (protein analysis).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge use of the services and facilities of the Center for Shared Use of Scientific Equipment of Ivanovo State University of Chemistry and Technology (with the support of the Ministry of Science and Higher Education of the Russian Federation grant number 075-15-2021-671), the Centre for Joint Use of Scientific Equipment ‘The Upper Volga Region Centre of Physico-Chemical Research’.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lavis, L.D.; Raines, R.T. Bright Ideas for Chemical Biology. ACS Chem. Biol. 2008, 3, 142–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Specht, E.A.; Braselmann, E.; Palmer, A.E. A Critical and Comparative Review of Fluorescent Tools for Live-Cell Imaging. Annu. Rev. Physiol. 2017, 79, 93–117. [Google Scholar] [CrossRef] [PubMed]
  3. Kaur, P.; Singh, K. Recent Advances in the Application of BODIPY in Bioimaging and Chemosensing. J. Mater. Chem. C 2019, 7, 11361–11405. [Google Scholar] [CrossRef]
  4. Nakamura, M.; Tsumoto, K.; Ishimura, K.; Kumagai, I. Detection of Biotinylated Proteins in Polyacrylamide Gels Using an Avidin-Fluorescein Conjugate. Anal. Biochem. 2002, 304, 231–235. [Google Scholar] [CrossRef] [PubMed]
  5. Hama, Y.; Urano, Y.; Koyama, Y.; Kamiya, M.; Bernardo, M.; Paik, R.S.; Shin, I.S.; Paik, C.H.; Choyke, P.L.; Kobayashi, H. A Target Cell-Specific Activatable Fluorescence Probe for In Vivo Molecular Imaging of Cancer Based on a Self-Quenched Avidin-Rhodamine Conjugate. Cancer Res. 2007, 67, 2791–2799. [Google Scholar] [CrossRef] [Green Version]
  6. Beija, M.; Afonso, C.A.M.; Martinho, J.M.G. Synthesis and Applications of Rhodamine Derivatives as Fluorescent Probes. Chem. Soc. Rev. 2009, 38, 2410–2433. [Google Scholar] [CrossRef] [Green Version]
  7. Ferreira, S.Z.; Carneiro, H.C.; Lara, H.A.; Alves, R.B.; Resende, J.M.; Oliveira, H.M.; Silva, L.M.; Santos, D.A.; Freitas, R.P. Synthesis of a New Peptide-Coumarin Conjugate: A Potential Agent against Cryptococcosis. ACS Med. Chem. Lett. 2015, 6, 271–275. [Google Scholar] [CrossRef] [Green Version]
  8. Ni, Y.; Wu, J. Far-Red and Near Infrared BODIPY Dyes: Synthesis and Applications for Fluorescent pH Probes and Bio-Imaging. Org. Biomol. Chem. 2014, 12, 3774–3791. [Google Scholar] [CrossRef]
  9. Kowada, T.; Maeda, H.; Kikuchi, K. BODIPY-Based Probes for the Fluorescence Imaging of Biomolecules in Living Cells. Chem. Soc. Rev. 2015, 44, 4953–4972. [Google Scholar] [CrossRef]
  10. Antina, E.; Bumagina, N.; Marfin, Y.; Guseva, G.; Nikitina, L.; Sbytov, D.; Telegin, F. BODIPY Conjugates as Functional Compounds for Medical Diagnostics and Treatment. Molecules 2022, 27, 1396. [Google Scholar] [CrossRef]
  11. Marek, M.; Kaiser, K.; Gruber, H.J. Biotin-Pyrene Conjugates with Poly(Ethylene Glycol) Spacers are Convenient Fluorescent Probes for Avidin and Streptavidin. Bioconjug. Chem. 1997, 8, 560–566. [Google Scholar] [CrossRef] [PubMed]
  12. Carreon, J.R.; Stewart, K.M.; Mahon, K.P.; Shin, S.; Kelley, S.O. Cyanine Dye Conjugates as Probes for Live Cell Imaging. Bioorg. Med. Chem. Lett. 2007, 17, 5182–5185. [Google Scholar] [CrossRef]
  13. Loudet, A.; Burgess, K. BODIPY Dyes and Their Derivatives: Syntheses and Spectroscopic Properties. Chem. Rev. 2007, 107, 4891–4932. [Google Scholar] [CrossRef] [PubMed]
  14. Boens, N.; Leen, V.; Dehaen, W. Fluorescent Indicators Based on BODIPY. Chem. Soc. Rev. 2012, 41, 1130–1172. [Google Scholar] [CrossRef] [PubMed]
  15. Rezende, L.C.D.; Emery, F.S. A Review of the Synthetic Strategies for the Development of BODIPY Dyes for Conjugation with Proteins. Orbital Electron. J. Chem. 2013, 5, 62–83. [Google Scholar] [CrossRef]
  16. Dilek, O.; Bane, S.L. Synthesis, Spectroscopic Properties and Protein Labeling of Water Soluble 3,5-Disubstituted Boron Dipyrromethenes. Bioorg. Med. Chem. Lett. 2009, 19, 6911–6913. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, D.; Fan, J.; Gao, X.; Wang, B.; Sun, S.; Peng, X. Carboxyl BODIPY Dyes from Bicarboxylic Anhydrides: One-Pot Preparation, Spectral Properties, Photostability, and Biolabeling. J. Org. Chem. 2009, 74, 7675–7683. [Google Scholar] [CrossRef]
  18. Poirel, A.; Retailleau, P.; de Nicola, A.; Ziessel, R. Synthesis of Water-Soluble Red-Emitting Thienyl-BODIPYs and Bovine Serum Albumin Labeling. Chem. Eur. J. 2014, 20, 1252–1257. [Google Scholar] [CrossRef]
  19. Kim, D.; Ma, D.; Kim, M.; Jung, Y.; Kim, N.H.; Lee, C.; Cho, S.W.; Park, S.; Huh, Y.; Jung, J.; et al. Fluorescent Labeling of Protein Using Blue-Emitting 8-Amino-BODIPY Derivatives. J. Fluoresc. 2017, 27, 2231–2238. [Google Scholar] [CrossRef]
  20. Ksenofontova, K.V.; Ksenofontov, A.A.; Khodov, I.A.; Rumyantsev, E.V. Novel BODIPY-Conjugated Amino Acids: Synthesis and Spectral Properties. J. Mol. Liq. 2019, 283, 695–703. [Google Scholar] [CrossRef]
  21. Amorim, V.G.; Melo, S.M.G.; Leite, R.F.; Coutinho, P.A.; da Silva, S.M.P.; Silva, A.R.; Amorim, F.G.; Pires, R.G.W.; Coitinho, J.B.; Emery, F.S.; et al. Synthesis and Characterization of Two Novel Red-Shifted Isothiocyanate BODIPYs and Their Application in Protein Conjugation. Dyes Pigment. 2020, 182, 108646:1–108646:7. [Google Scholar] [CrossRef]
  22. Jeon, S.; Kim, T.-I.; Jin, H.; Lee, U.; Bae, J.; Bouffard, J.; Kim, Y. Amine-Reactive Activated Esters of meso-CarboxyBODIPY: Fluorogenic Assays and Labeling of Amines, Amino Acids, and Proteins. J. Am. Chem. Soc. 2020, 142, 9231–9239. [Google Scholar] [CrossRef] [PubMed]
  23. Ksenofontova, K.V.; Ksenofontov, A.A.; Khodov, I.A.; Rumyantsev, E.V. Synthesis and Study of Spectral Properties of Amino Acids-BODIPY Conjugates. Izv. Vyssh. Uchebn. Zaved. Khim. Khim. Tekhnol. 2020, 63, 4–11. [Google Scholar] [CrossRef]
  24. Raskolupova, V.I.; Popova, T.V.; Zakharova, O.D.; Nikotina, A.E.; Abramova, T.V.; Silnikov, V.N. Human Serum Albumin Labelling with a New BODIPY Dye Having a Large Stokes Shift. Molecules 2021, 26, 2679. [Google Scholar] [CrossRef]
  25. Brellier, M.; Duportail, G.; Baati, R. Convenient Synthesis of Water-Soluble Nitrilotriacetic Acid (NTA) BODIPY Dyes. Tetrahedron Lett. 2010, 51, 1269–1272. [Google Scholar] [CrossRef]
  26. Lakowicz, J.R. Principles of Fluorescence Spectroscopy, 3rd ed.; Springer: New York, NY, USA, 2006; ISBN 978-0-387-46312-4. [Google Scholar]
  27. Kamlet, M.J.; Abboud, J.L.M.; Abraham, M.H.; Taft, R.W. Linear Solvation Energy Relationships. 23. A Comprehensive Collection of the Solvatochromic Parameters, π*, α, and β, and Some Methods for Simplifying the Generalized Solvatochromic Equation. J. Org. Chem. 1983, 48, 2877–2887. [Google Scholar] [CrossRef]
  28. Catalán, J. Toward a Generalized Treatment of the Solvent Effect Based on Four Empirical Scales: Dipolarity (SdP, a New Scale), Polarizability (SP), Acidity (SA), and Basicity (SB) of the Medium. J. Phys. Chem. B 2009, 113, 5951–5960. [Google Scholar] [CrossRef]
  29. Mataga, N.; Kaifu, Y.; Koizumi, M. Solvent Effects upon Fluorescence Spectra and the Dipolemoments of Excited Molecules. Bull. Chem. Soc. Jpn. 1956, 29, 465–470. [Google Scholar] [CrossRef] [Green Version]
  30. Boens, N.; Leen, V.; Dehaen, W.; Wang, L.; Robeyns, K.; Qin, W.; Tang, X.; Beljonne, D.; Tonnelé, C.; Paredes, J.M.; et al. Visible Absorption and Fluorescence Spectroscopy of Conformationally Constrained, Annulated BODIPY Dyes. J. Phys. Chem. A 2012, 116, 9621–9631. [Google Scholar] [CrossRef]
  31. Cieślik-Boczula, K.; Burgess, K.; Li, L.; Nguyen, B.; Pandey, L.; de Borggraeve, W.M.; van der Auweraer, M.; Boens, N. Photophysics and Stability of Cyano-Substituted Boradiazaindacene Dyes. Photochem. Photobiol. Sci. 2009, 8, 1006–1015. [Google Scholar] [CrossRef]
  32. Banerjee, S.; Stüker, T.; Saalfrank, P. Vibrationally Resolved Optical Spectra of Modified Diamondoids Obtained from Time-Dependent Correlation Function Methods. Phys. Chem. Chem. Phys. 2015, 17, 19656–19669. [Google Scholar] [CrossRef] [Green Version]
  33. McWeeny, R. Some Recent Advances in Density Matrix Theory. Rev. Mod. Phys. 1960, 32, 335–369. [Google Scholar] [CrossRef]
  34. Tariq, R.; Khera, R.A.; Rafique, H.; Azeem, U.; Naveed, A.; Ayub, A.R.; Iqbal, J. Computational and Theoretical Study of Subphthalocyanine Based Derivatives by Varying Acceptors to Increase the Efficiency of Organic Solar Cells. Comput. Theor. Chem. 2021, 1203, 113356:1–113356:11. [Google Scholar] [CrossRef]
  35. Tretiak, S.; Igumenshchev, K.; Chernyak, V. Exciton Sizes of Conducting Polymers Predicted by Time-Dependent Density Functional Theory. Chem. Rev. 2005, 71, 3171. [Google Scholar] [CrossRef] [Green Version]
  36. Titov, E. On the Low-Lying Electronically Excited States of Azobenzene Dimers: Transition Density Matrix Analysis. Molecules 2021, 26, 4245. [Google Scholar] [CrossRef]
  37. Hermanson, G.T. Bioconjugate Techniques, 3rd ed.; Elsevier: London, UK, 2013; ISBN 978-0-12-382239-0. [Google Scholar]
  38. Jones Brunette, A.M.; Farrens, D.L. Distance Mapping in Proteins Using Fluorescence Spectroscopy: Tyrosine, Like Tryptophan, Quenches Bimane Fluorescence in a Distance-Dependent Manner. Biochemistry 2014, 53, 6290–6301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Karolin, J.; Johansson, L.B.-A.; Strandberg, L.; Ny, T. Fluorescence and Absorption Spectroscopic Properties of Dipyrrometheneboron Difluoride (BODIPY) Derivatives in Liquids, Lipid Membranes, and Proteins. J. Am. Chem. Soc. 1994, 116, 7801–7806. [Google Scholar] [CrossRef]
  40. Wang, Y.-Q.; Zhang, H.-M.; Zhang, G.-C. Studies of the Interaction between Palmatine Hydrochloride and Human Serum Albumin by Fluorescence Quenching Method. J. Pharm. Biomed. Anal. 2006, 41, 1041–1046. [Google Scholar] [CrossRef]
  41. Ramezani, F.; Rafii-Tabar, H. An In-Depth View of Human Serum Albumin Corona on Gold Nanoparticles. Mol. Biosyst. 2015, 11, 454–462. [Google Scholar] [CrossRef]
  42. Zsila, F. Subdomain IB is the Third Major Drug Binding Region of Human Serum Albumin: Toward the Three-Sites Model. Mol. Pharm. 2013, 10, 1668–1682. [Google Scholar] [CrossRef]
  43. Fanali, G.; Di Masi, A.; Trezza, V.; Marino, M.; Fasano, M.; Ascenzi, P. Human Serum Albumin: From Bench to Bedside. Mol. Asp. Med. 2012, 33, 209–290. [Google Scholar] [CrossRef] [PubMed]
  44. Iskhakova, Z.I.; Zhuravleva, D.E.; Heim, C.; Hartmann, M.D.; Laykov, A.V.; Forchhammer, K.; Kayumov, A.R. PotN Represents a Novel Energy-State Sensing PII Subfamily, Occurring in Firmicutes. FEBS J. 2022, 289, 5305–5321. [Google Scholar] [CrossRef] [PubMed]
  45. Heinrich, A.; Woyda, K.; Brauburger, K.; Meiss, G.; Detsch, C.; Stülke, J.; Forchhammer, K. Interaction of the Membrane-Bound GlnK-AmtB Complex with the Master Regulator of Nitrogen Metabolism TnrA in Bacillus subtilis. J. Biol. Chem. 2006, 281, 34909–34917. [Google Scholar] [CrossRef] [Green Version]
  46. McPhie, P. [4] Dialysis. In Enzyme Purification and Related Techniques; Jakoby, W.B., Ed.; Elsevier: Amsterdam, The Netherlands, 1971; pp. 23–32. ISBN 978-0-12-181885-2. [Google Scholar]
  47. Laemmli, U.K. Cleavage of Structural Proteins during the Assembly of the Head of Bacteriophage T4. Nature 1970, 227, 680–685. [Google Scholar] [CrossRef] [PubMed]
  48. Jameson, L.P.; Dzyuba, S.V. Expeditious, Mechanochemical Synthesis of BODIPY Dyes. Beilstein J. Org. Chem. 2013, 9, 786–790. [Google Scholar] [CrossRef]
  49. Haugland, R.P. Coupling of Monoclonal Antibodies with Fluorophores. In Monoclonal Antibody Protocols; Davis, W.C., Ed.; Humana Press: Totowa, NJ, USA, 1995; pp. 205–221. ISBN 978-1-59259-532-7. [Google Scholar]
  50. Würth, C.; Grabolle, M.; Pauli, J.; Spieles, M.; Resch-Genger, U. Relative and Absolute Determination of Fluorescence Quantum Yields of Transparent Samples. Nat. Protoc. 2013, 8, 1535–1550. [Google Scholar] [CrossRef]
  51. Bannwarth, C.; Caldeweyher, E.; Ehlert, S.; Hansen, A.; Pracht, P.; Seibert, J.; Spicher, S.; Grimme, S. Extended Tight-Binding Quantum Chemistry Methods. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2021, 11, e1494:1–e1494:49. [Google Scholar] [CrossRef]
  52. Pracht, P.; Bohle, F.; Grimme, S. Automated Exploration of the Low-Energy Chemical Space with Fast Quantum Chemical Methods. Phys. Chem. Chem. Phys. 2020, 22, 7169–7192. [Google Scholar] [CrossRef]
  53. Wang, J.; Durbeej, B. How Accurate are TD-DFT Excited-State Geometries Compared to DFT Ground-State Geometries? J. Comput. Chem. 2020, 41, 1718–1729. [Google Scholar] [CrossRef]
  54. Santoro, F.; Cerezo, J.; FCclasses3. A Code for Vibronic Calculations. Available online: http://www.iccom.cnr.it/en/fcclasses (accessed on 12 October 2022).
  55. Neese, F. Software Update: The ORCA Program System-Version 5.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2022, 12, e1606:1–e1606:15. [Google Scholar] [CrossRef]
  56. Neese, F. Software Update: The ORCA Program System, Version 4.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8, e1327:1–e1327:6. [Google Scholar] [CrossRef]
  57. Neese, F. The ORCA Program System. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  58. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  59. Chemcraft-Graphical Software for Visualization of Quantum Chemistry Computations. Available online: https://www.chemcraftprog.com (accessed on 12 October 2022).
  60. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  61. Morris, G.M.; Huey, R.; Lindstrom, W.; Sanner, M.F.; Belew, R.K.; Goodsell, D.S.; Olson, A.J. AutoDock4 and AutoDockTools4: Automated Docking with Selective Receptor Flexibility. J. Comput. Chem. 2009, 30, 2785–2791. [Google Scholar] [CrossRef] [Green Version]
  62. Protein Data Bank. 4F5S: Crystal Structure of Bovine Serum Albumin. Available online: https://www.rcsb.org/structure/4F5S (accessed on 20 August 2022).
  63. Protein Data Bank. 7O4X: Crystal Structure of the PII-Like Protein PotN from Lentilactobacillus Hilgardii. Available online: https://www.rcsb.org/structure/7O4X (accessed on 20 August 2022).
  64. Morris, G.M.; Goodsell, D.S.; Halliday, R.S.; Huey, R.; Hart, W.E.; Belew, R.K.; Olson, A.J. Automated Docking Using a Lamarckian Genetic Algorithm and an Empirical Binding Free Energy Function. J. Comput. Chem. 1998, 19, 1639–1662. [Google Scholar] [CrossRef]
  65. Zhu, K.; Borrelli, K.W.; Greenwood, J.R.; Day, T.; Abel, R.; Farid, R.S.; Harder, E. Docking Covalent Inhibitors: A Parameter Free Approach to Pose Prediction and Scoring. J. Chem. Inf. Model. 2014, 54, 1932–1940. [Google Scholar] [CrossRef]
  66. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A Visualization System for Exploratory Research and Analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef]
Scheme 1. Scheme of reactive BODIPY synthesis.
Scheme 1. Scheme of reactive BODIPY synthesis.
Molecules 27 07911 sch001
Figure 1. Absorption (unfilled area plot) and emission (filled area plot) spectra of COOH-Ph-BODIPY (blue) and NHS-Ph-BODIPY (green) in DMSO.
Figure 1. Absorption (unfilled area plot) and emission (filled area plot) spectra of COOH-Ph-BODIPY (blue) and NHS-Ph-BODIPY (green) in DMSO.
Molecules 27 07911 g001
Figure 2. Relationships between experimental and calculated maximum absorption νabs(max) (open circle) and emission νem(max) (full circle) wavenumbers of COOH-Ph-BODIPY (a) and NHS-Ph-BODIPY (b) obtained by multiple linear regression analysis according to Equation (5) in which y0, ad are regression coefficients (numbers refer to organic solvents in Table 1).
Figure 2. Relationships between experimental and calculated maximum absorption νabs(max) (open circle) and emission νem(max) (full circle) wavenumbers of COOH-Ph-BODIPY (a) and NHS-Ph-BODIPY (b) obtained by multiple linear regression analysis according to Equation (5) in which y0, ad are regression coefficients (numbers refer to organic solvents in Table 1).
Molecules 27 07911 g002
Figure 3. Ground S0 (blue) and first excited S1 (red) molecular geometry comparison for COOH-Ph-BODIPY (a) and NHS-Ph-BODIPY (b).
Figure 3. Ground S0 (blue) and first excited S1 (red) molecular geometry comparison for COOH-Ph-BODIPY (a) and NHS-Ph-BODIPY (b).
Molecules 27 07911 g003
Figure 4. Theoretical vibrationally resolved absorption (unfilled area plot) and emission (filled area plot) spectra of COOH-Ph-BODIPY (blue) and NHS-Ph-BODIPY (green) in DMSO.
Figure 4. Theoretical vibrationally resolved absorption (unfilled area plot) and emission (filled area plot) spectra of COOH-Ph-BODIPY (blue) and NHS-Ph-BODIPY (green) in DMSO.
Molecules 27 07911 g004
Figure 5. Partitioned transition density for first singlet excitation of COOH-Ph-BODIPY (b) and NHS-Ph-BODIPY (c) in DMSO and corresponding molecular segments (a).
Figure 5. Partitioned transition density for first singlet excitation of COOH-Ph-BODIPY (b) and NHS-Ph-BODIPY (c) in DMSO and corresponding molecular segments (a).
Molecules 27 07911 g005
Scheme 2. Scheme of fluorescent labeling of proteins.
Scheme 2. Scheme of fluorescent labeling of proteins.
Molecules 27 07911 sch002
Figure 6. Absorption (unfilled area plot) and emission (filled area plot) spectra of NHS-Ph-BODIPY (green) and BSA–NHS-Ph-BODIPY (orange) in a DMSO–bicarbonate buffer (pH 8.3) mixture (1:9).
Figure 6. Absorption (unfilled area plot) and emission (filled area plot) spectra of NHS-Ph-BODIPY (green) and BSA–NHS-Ph-BODIPY (orange) in a DMSO–bicarbonate buffer (pH 8.3) mixture (1:9).
Molecules 27 07911 g006
Figure 7. SDS-PAGE of double dilutions of BSA (MW = 66.2 kDa) (left) and PotN (MW = 13.7 kDa) (right) stained with Coomassie brilliant blue (a,c) and labeled with NHS-Ph-BODIPY (b,d) (each gel contains protein weight marker and unlabeled protein).
Figure 7. SDS-PAGE of double dilutions of BSA (MW = 66.2 kDa) (left) and PotN (MW = 13.7 kDa) (right) stained with Coomassie brilliant blue (a,c) and labeled with NHS-Ph-BODIPY (b,d) (each gel contains protein weight marker and unlabeled protein).
Molecules 27 07911 g007
Figure 8. Basic native-PAGE of NHS-Ph-BODIPY-labeled (+ Flu) and unlabeled PotN (MW = 13.7 kDa, pI = 6.6), PotAc (MW = 18.1 kDa, pI = 5.9), GS (MW = 50.3 kDa, pI = 5.1), and BSA (MW = 66.2 kDa, pI = 5.9) visualized with Coomassie brilliant blue (a) and using UV transillumination (b) (each lane contains 10 µg of protein).
Figure 8. Basic native-PAGE of NHS-Ph-BODIPY-labeled (+ Flu) and unlabeled PotN (MW = 13.7 kDa, pI = 6.6), PotAc (MW = 18.1 kDa, pI = 5.9), GS (MW = 50.3 kDa, pI = 5.1), and BSA (MW = 66.2 kDa, pI = 5.9) visualized with Coomassie brilliant blue (a) and using UV transillumination (b) (each lane contains 10 µg of protein).
Molecules 27 07911 g008
Figure 9. Binding site of NHS-Ph-BODIPY in BSA (a) and its amino acid composition (b) according to blind docking results.
Figure 9. Binding site of NHS-Ph-BODIPY in BSA (a) and its amino acid composition (b) according to blind docking results.
Molecules 27 07911 g009
Figure 10. Amino acid environment of NHS-Ph-BODIPY in BSA according to covalent docking results of the Lys221-BSA–NHS-Ph-BODIPY system.
Figure 10. Amino acid environment of NHS-Ph-BODIPY in BSA according to covalent docking results of the Lys221-BSA–NHS-Ph-BODIPY system.
Molecules 27 07911 g010
Figure 11. Binding site of NHS-Ph-BODIPY in PotN (a) and its amino acid composition (b) according to blind docking results.
Figure 11. Binding site of NHS-Ph-BODIPY in PotN (a) and its amino acid composition (b) according to blind docking results.
Molecules 27 07911 g011
Table 1. Photophysical characteristics of COOH-Ph-BODIPY and NHS-Ph-BODIPY in organic solvents.
Table 1. Photophysical characteristics of COOH-Ph-BODIPY and NHS-Ph-BODIPY in organic solvents.
Solventλabs max, nmλem max, nmΔν, cm−1ε, M−1·cm−1Φτ, nskr, ns−1knr, ns−1
COOH-Ph-BODIPY
1n-Hexane52754355998,5000.613.230.190.12
2Toluene52954968999,5000.614.090.150.09
3Ethyl acetate524541600118,5000.734.270.170.06
4Chloroform52954658991,4000.564.320.130.10
5DCM52854765861,5000.383.760.100.16
6Acetone52354267051,9000.323.66*0.090.19
7DMSO52554156370,9000.444.570.090.12
8Acetonitrile52153860659,2000.364.25*0.090.15
9n-Propanol52353956870,4000.434.76*0.090.12
10Ethanol52353956856,7000.354.85*0.070.13
NHS-Ph-BODIPY
1n-Hexane52954762269,8000.353.170.110.21
2Toluene53255268161,9000.434.020.110.14
3Ethyl acetate52754459356,9000.563.700.150.12
4Chloroform53155065173,1000.463.640.130.15
5DCM53055068655,0000.403.750.110.16
6Acetone52654566355,2000.353.230.110.20
7DMSO52954865559,6000.474.140.110.13
8Acetonitrile52554466546,5000.393.430.110.18
9n-Propanol52754666052,6000.493.600.130.14
10Ethanol52754562754,0000.383.850.100.16
Table 2. Regression coefficients y0, ad and coefficients of determination R2 for multiple linear regression analysis of maximum absorption νabs(max) and emission νem(max) wavenumbers and Stokes shifts Δν of COOH-Ph-BODIPY and NHS-Ph-BODIPY as a function of Catalán solvent scale parameters {SA, SB, SP, SdP}.
Table 2. Regression coefficients y0, ad and coefficients of determination R2 for multiple linear regression analysis of maximum absorption νabs(max) and emission νem(max) wavenumbers and Stokes shifts Δν of COOH-Ph-BODIPY and NHS-Ph-BODIPY as a function of Catalán solvent scale parameters {SA, SB, SP, SdP}.
y0aSAbSBcSPdSdPR2
COOH-Ph-BODIPY
νabs(max), cm−119,553 ± 129−108 ± 113131 ± 77−988 ± 190203 ± 540.93
νem(max), cm−118,901 ± 29113 ± 254184 ± 172−924 ± 482157 ± 1220.75
Δν, cm−1653 ± 189−122 ± 165−53 ± 112−64 ± 27846 ± 790.29
NHS-Ph-BODIPY
νabs(max), cm−119,461 ± 76−167 ± 66100 ± 45−940 ± 112150 ± 320.96
νem(max), cm−118,912 ± 104−258 ± 91204 ± 62−1064 ± 15481 ± 440.94
Δν, cm−1549 ± 7391 ± 63−104 ± 43124 ± 10768 ± 310.70
Table 3. Ground S0 and first excited S1 total dipole moments ||µGS|| and ||µTD|| of COOH-Ph-BODIPY and NHS-Ph-BODIPY, difference between their absolute values, and absolute value of their difference in n-hexane and DMSO.
Table 3. Ground S0 and first excited S1 total dipole moments ||µGS|| and ||µTD|| of COOH-Ph-BODIPY and NHS-Ph-BODIPY, difference between their absolute values, and absolute value of their difference in n-hexane and DMSO.
CompoundSolvent||µGS||, a.u.||µTD||, a.u.||µGS|| − ||µTD||, a.u.||µGSµTD||, a.u.
COOH-Ph-BODIPYn-Hexane1.4050.7240.6810.714
DMSO1.6580.9750.6830.711
NHS-Ph-BODIPYn-Hexane1.9901.3100.6790.718
DMSO2.3361.6590.6780.711
Table 4. Photophysical characteristics of NHS-Ph-BODIPY and BSA–NHS-Ph-BODIPY in a DMSO–bicarbonate buffer (pH 8.3) mixture (1:9).
Table 4. Photophysical characteristics of NHS-Ph-BODIPY and BSA–NHS-Ph-BODIPY in a DMSO–bicarbonate buffer (pH 8.3) mixture (1:9).
λabs max, nmλem max, nmΔν, cm−1Φτ1, nsτ2, ns
NHS-Ph-BODIPY5265425610.414.96
BSA–NHS-Ph-BODIPY28035071430.302.798.00
530554817
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ksenofontova, K.V.; Kerner, A.A.; Ksenofontov, A.A.; Shagurin, A.Y.; Bocharov, P.S.; Lukanov, M.M.; Kayumov, A.R.; Zhuravleva, D.E.; Iskhakova, Z.I.; Molchanov, E.E.; et al. Amine-Reactive BODIPY Dye: Spectral Properties and Application for Protein Labeling. Molecules 2022, 27, 7911. https://doi.org/10.3390/molecules27227911

AMA Style

Ksenofontova KV, Kerner AA, Ksenofontov AA, Shagurin AY, Bocharov PS, Lukanov MM, Kayumov AR, Zhuravleva DE, Iskhakova ZI, Molchanov EE, et al. Amine-Reactive BODIPY Dye: Spectral Properties and Application for Protein Labeling. Molecules. 2022; 27(22):7911. https://doi.org/10.3390/molecules27227911

Chicago/Turabian Style

Ksenofontova, Ksenia V., Anastasia A. Kerner, Alexander A. Ksenofontov, Artyom Yu. Shagurin, Pavel S. Bocharov, Michael M. Lukanov, Airat R. Kayumov, Darya E. Zhuravleva, Zalina I. Iskhakova, Evgeniy E. Molchanov, and et al. 2022. "Amine-Reactive BODIPY Dye: Spectral Properties and Application for Protein Labeling" Molecules 27, no. 22: 7911. https://doi.org/10.3390/molecules27227911

Article Metrics

Back to TopTop