Isomerism and coordination mode effects on two-photon absorption of tris(picolyl)amine-based fluorescent probes for zinc ions
Zhao Ke, Song Jun, Zhu Mei-Yu, Zhang Han, Wang Chuan-Kui
School of Physics and Electronics, Shandong Normal University, Jinan 250358, China

 

† Corresponding author. E-mail: zhaoke@sdnu.edu.cn

Project supported by the Natural Science Foundation of Shandong Province, China (Grant No. ZR2014AM026) and the Taishan Scholar Project of Shandong Province, China.

Abstract

One-photon absorption and two-photon absorption (TPA) properties of three tris(picolyl)amine-based zinc ion sensors are investigated by employing the density functional response theory in combination with the polarizable continuum model. The different isomer and coordination geometry of each probe are taken into account. Special emphasis is placed on the effects of isomerism and the coordination mode on the optical properties. The intra-molecular charge transfer (ICT) properties are specified by natural bond orbital charge analysis. It is shown that the isomerism has non-negligible effects on TPA properties of free ligands. It is found that both the TPA wavelength and the cross section are highly dependent on the coordination mode. When the zinc ion connects with the picolyl unit in the middle of a ligand, the zinc complex has a large TPA intensity in a long wavelength range due to the increased ICT mechanism.

1. Introduction

Great progress has been made in the imaging of living cells and tissues since the first two-photon fluorescence microscopy (TPM) appeared in 1990.[1] Compared with one-photon microscopy, TPM, in which two near-infrared photons for excitation are used, offers several advantages, including increased penetration depth, lower tissue auto-fluorescence and self-absorption, as well as reduced photo-damage and photo-bleaching.[24] As such, to facilitate the utility of TPM in bioimaging applications, there is a strong need to design two-photon (TP) probes for specific applications. To date, a variety of TP probes for diverse analytes, such as cation and anion ions,[58] pH values,[9,10] small molecules,[11,12] and DNA,[13] have been developed and their performances in bioimaging applications have been explored.

Zinc ion is the second most abundant transition metal ion in the living body. It plays an important role in many biological processes.[1416] The dysbolism of zinc ions will result in many human diseases such as dysimmunity, Alzheimer’s, and Parkinson’s diseases. Therefore, it is necessary to detect the spatiotemporal distributions of the zinc ions in living systems. Recently, a number of zinc ion TP probes have been synthesized and their bioimaging applications have been demonstrated.[5,6] However, the related theoretical study is still very limited.[1721] Wang et al.[17] designed a series of TP probes for zinc ions based on the intra-molecular charge transfer (ICT) mechanism. The one-photon absorption (OPA), two-photon absorption (TPA), and fluorescence properties were calculated. It was found that the designed probes can become fluorescent bioimaging reagents used for ratiometric detection. Bednarska et al.[18] investigated the responsive mechanism for a bipyridine-centered ratiometric zinc ions probe theoretically. The electric structure and nonlinear absorption spectra were calculated including solute-solvent interactions. Their results indicated that nonlinear response functions combined with density functional theory (DFT) can be successful in analyzing metal ion probes. In our previous researches,[1921] response theory was used to investigate the OPA and TPA of a series of zinc ion sensors. The responsive mechanisms were specified at length, and a series of TP probes was designed by taking the established structure-property relationships into account.

In this paper, we choose three zinc ion TP probes 1, 2, and 3 (see Fig. 1) as the model molecules. It was reported that probe 1 undergoes a significant increase of the TPA cross sections by the Zn2+-coordination induced ICT enhancement.[22] To explore the recognition mechanisms of the probes and the structure-property relationships, we perform a theoretical study on the OPA and TPA properties of the probes before and after combination with Zn2+ using DFT. In our study, it is interesting to find that these probes could have several isomeric geometries and then the corresponding zinc complexes pose various coordination modes. Therefore, we put an emphasis on the effects of isomerism and coordination mode on the TPA properties for the probe isomers and their zinc complexes.

Fig. 1. Chemical structures of probes 1, 2, and 3.

Although isomerism is quite a common phenomenon in chemistry, the effects of isomerism on TPA properties of organic molecules have seldom been discussed in the literature.[23,24] In our previous researches,[2529] we investigated the influences of isomerism on TPA properties for a series of TPA active molecules. For a probe in solution at room temperature, various isomers of free and bound fluorophores are possible and produce specific spectral features. It is therefore highly required to take isomerism and coordination mode effects into account. To the best of our knowledge, these effects have not been investigated theoretically for TP metal ion sensors. Our research will provide a good understanding of the experimental observations for TP metal ion probes.

2. Computational method

The oscillator strength can be used to describe the intensity of OPA. It is expressed as:

where α ∈ (x,y,z), μα is the dipole moment operator, and ωf denotes the excitation energy from the ground state 〈0| to the excited state |f〉.

In the case of resonant degenerate TPA, the sum-over-state expression for the two-photon matrix element can be written as[30]

where α,β ∈ (x,y,z), and ω is the fundamental frequency of the laser beam, and we assume that it is equal to half the excitation energy of the final state, i.e., 2ω = ωf; ωs represents the excitation energy for the intermediate state |s〉. The summation here includes all intermediate, initial, and final states. In response theory, the two-photon matrix element Sα β can be calculated through the single residues of the quadratic response function.[31]

The microscopic TPA cross section of molecules excited by a linear polarized single beam can be expressed as[30,32]

The macroscopic TPA cross section that can be directly compared with the experimental measurement, is defined as[32]

Here a0 is the Bohr radius, c is the speed of light, and α is the fine structure constant. The level broadening Γ of the final state is assumed to have the commonly used value Γ = 0.1 eV. The TPA cross section is in units of GM, 1 GM = 10−50 cm4°s/photon.

In this work, the geometries of molecules are fully optimized by using DFT with the 6-31G(d,p) basis set and the B3LYP hybrid functional. On the basis of the optimized structures, the OPA properties are computed by the time-dependent DFT (TD-DFT) approach at the B3LYP level with the 6-31G(d) basis set. All the calculations of optimization and OPA properties are carried out in the Gaussian 09 program.[33] The TPA cross sections are obtained by the response theory using the B3LYP functional with 6-31G(d) basis set in the Dalton 2013 package.[34] In addition, the effect of water solvent is taken into account through the self-consistent reaction field theory based on the polarizable continuum model (PCM) in both Gaussian and Dalton calculations. Our previous work has shown that the B3LYP functional calculations can give reasonable TPA properties which are consistent with the trend of experimental observations.[1929] The use of bigger basis sets could probably provide better numerical results, but we believe that the overall picture would not change.

3. Results and discussion
3.1. Geometry optimization

Probe 1 as depicted in Fig. 1 is composed of an electron-donating anisole substituent and an electron-deficient tris(picolyl)amine Zn2+ receptor linked through an oxazole bridge. The tris(picolyl)amine unit often acts as the Zn2+ receptor in the TP probe. The nitrogen atom could coordinate with the zinc ion by donating a lone electron pair into an empty metal orbital. In probe 2, we replace the anisole donor with the dimethylamino terminal because the dimethylamino group has stronger ability to donate and transport electrons. In probe 3, the hydroxy group with weaker donation ability is used. These push-pull structures can improve the ICT and induce a large TPA response. The geometry optimization of probe 1 gives as many as six equilibrium conformations 1a--1f as shown in Fig. 2. The frequency calculations for these geometries do not produce any imaginary frequencies. It can be seen that the six isomers differ by the orientations of two picolyl groups at the end of the ligands. The optimized geometries of zinc complexes are also presented in Fig. 2. Two Cl ions are incorporated explicitly in order to include counter ion effects and give the charge-neutral complex.[19,22] It is interesting to find that different coordination geometries are obtained. In Zn1a structure, the zinc ion prefers to chelate with four nitrogen atoms of the tris(picolyl)amine receptor to complete the six-coordination geometry. The bond length between the nitrogen atom and zinc ion has a typical value of 2.1 Å–2.3 Å, which is similar to the general x-ray crystal structure.[35] The Zn1b and Zn1c have similar structures, in which the zinc ion connects with the three nitrogen atoms of the dipicolylamine group at the end of the ligands. The Zn1d also has a five-coordination geometry, but one of the connecting nitrogen atoms belongs to the picolyl unit in the middle of the ligand. In Zn1e and Zn1f, only two nitrogen atoms take part in the coordination. In Zn1f, the zinc ion does not connect with the picolyl unit in the middle of the ligand. By examining these complexes' structures, one can find that the zinc ions in Zn1b, Zn1c, and Zn1f are a little farther away from the oxazole cores than the zinc ions in the other structures. It is also noticed that each of all the backbones for these structures has good planarity.

Fig. 2. (color online) Optimized geometries of 1a–1f and the corresponding zinc complexes.

The optimized geometries of probe 2 and the corresponding zinc complexes are displayed in Fig. 3. The result is the same as that of probe 1. Six isomers of probe 2 and six coordination conformations are also obtained. The corresponding geometries of probe 3 are also fully optimized.

Fig. 3. (color online) Optimized geometries of 2a–2f and the corresponding zinc complexes.

It is expected that the various conformations will give rise to different electronic structures, which are closely related to the TPA transition matrix element according to Eq. (2). The permanent dipole moments and the energies of these structures are examined and the results are listed in Table 1. It shows that the permanent dipole moments of isomers 1a–1f are quite different: 1c has the largest dipole moment (7.584 deb, 1 deb = 3.33564 × 10−30 C·m), and 1d has the smallest one (3.037 deb). In the corresponding zinc complexes, the permanent dipole moments still maintain almost the same order. This result is consistent with the characteristics of the structures. It is found that 1f has the lowest energy among six isomers of probe 1, and the energy differences between the other structures and 1f are very small in general. The energy difference ΔE ranges from 1.8 kcal/mol to 4.1 kcal/mol. Such small energy barriers indicate that these isomers could be thermally populated. It is surprising that Zn1a has the lowest energy among the zinc complexes and the energy difference between Zn1e and Zn1a has the largest value, which is 11.04 kcal/mol. This is probably because Zn1a has a six-coordination geometry while Zn1e only has four coordination bonds. It is also found that the energy differences between Zn1a and Zn1b and between Zn1a and Zn1d are only 0.667 kcal/mol and 0.733 kcal/mol, which means that these structures can be populated at considerable percentages according to the Boltzmann distribution. The case of probe 2 is similar to that of probe 1. The difference in permanent dipole moment among the isomers becomes more obvious. However, for probe 3, the difference in permanent dipole moment among the isomers is a little smaller than for probe 1.

Table 1.

Values of permanent dipole moment μ and energy difference ΔE.

.
3.2. One-photon absorption

Our calculated OPA wavelength λop, oscillator strength fop, and transition nature of all the studied molecules are listed in Table 2. It shows that the maximum OPA peaks of these compounds are all derived from the transition between ground state (S0) and the first excited state (S1), which is mainly dominated by the transition from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO).

For probe 1, one can see that the OPA wavelengths of the six isomers 1a–1f are all located at about 360 nm and their oscillator strengths only have a little difference. When the molecule binds to a zinc ion, the λop is red-shifted or blue-shifted to some extent. The Zn1a, Zn1d, Zn1e, and Zn1f have 16-, 21-, 28-, and 7-nm red-shifts respectively, while Zn1b and Zn1c have 4- and 3-nm blue-shift. The largest red-shift occurs in Zn1e. As a result, various coordination conformations have different OPA wavelengths. It also shows that the differences in oscillator strength among these zinc complexes increase in comparison with those among free ligands. The fop of Zn1e is 0.7, which is much lower than the value of Zn1f, 0.99. These results indicate that the coordination modes have some effects on OPA properties. One can conclude that when the zinc ion coordinates with the nitrogen atom, which belongs to the picolyl unit in the middle of the ligand, the OPA wavelength has a considerable red-shift.

From probe 1 to probe 2 it follows that the OPA wavelengths of 2a–2f are red-shifted to around 410 nm due to the stronger electron donor. The isomerism still has very little effect on the OPA properties of the free ligands. However, when it comes to the zinc complexes, the various coordination modes have quite different OPA properties. The OPA wavelengths of Zn2a, Zn2d, and Zn2e are red-shifted obviously with respect to their free ligands. The largest red-shift still occurs in Zn2e, arriving at 48 nm. In addition, the difference in λop between Zn2b and Zn2e becomes 47 nm.

Table 2.

One-photon absorption wavelength properties of the molecules in water solvent.

.

The OPA properties of probe 3 are similar to those of probe 1 due to the little difference in the donation ability between the methoxy and the hydroxy groups. Nearly all the OPA wavelengths of 3a–3f and Zn3a–Zn3f are slightly blue-shifted compared with those of the corresponding structures 1a–1f and Zn1a–Zn1f.

3.3. Two-photon absorption

The TPA wavelength λtp and cross section α in water solvent for all the studied molecules are calculated by the response theory in the Dalton program. The TPA cross sections of the three lowest excited states for probes 1 and 2 are presented in Table 3.

Table 3.

Values of two-photon absorption wavelengths λtp and cross section σ of the three lowest excited states of probes 1 and 2 in water solvent.

.

For 1a, 1d, and 1e, the first excited state has the largest TPA intensity. The largest cross sections of 1a, 1d, and 1e are located at about 730 nm with 224, 237, and 260 GM respectively. As for 1b and 1c, the second excited state has the largest TPA intensity and the corresponding TPA wavelength λtp is centered at around 700 nm. The values of σ are equal to 290 GM and 373 GM. For 1f, both the first and the second excited states have considerable TPA cross sections. We note that the isomerism effects cannot be neglected. If all the isomers are taken into account, the obtained TPA spectrum could have several separated absorption maxima at about 670 nm, 700 nm, and 730 nm respectively.

When coordinated with Zn2+, the first excited state has the largest TPA intensity for all the zinc complexes. One can see that the coordination mode has an important effect on the TPA property. There are big differences in not only TPA intensity but also wavelength among zinc complexes. The largest cross section of Zn1a is 367 GM at 773 nm. The TPA wavelengths of Zn1b and Zn1c are 736 nm and 745 nm respectively, which are much shorter than that of Zn1a. Zn1b has the smallest cross section, which is only 279 GM. The λtp of Zn1d is similar to that of Zn1a, but the σ increases to 403 GM. Zn1e has the longest λtp, 808 nm, and the largest σ, 439 GM. That is, the absorption position ranges from 736 nm to 808 nm, which means that the absorption bands have large shifts between each other. From the cross section, we find that the TPA intensity is in the order Zn1e>Zn1d>Zn1f>Zn1a>Zn1c>Zn1b. The difference between the maximum and the minimum cross sections is large, which is 160 GM, about 36% of the maximum. By the inspection of the structures of the zinc complexes, one can quickly realize that the coordination mode with the connection between the zinc ion and the nitrogen atoms, which belongs to the picolyl unit in the middle of the ligand, always brings about the large cross sections in a long wavelength region. According to Eq. (2), we know that the TPA transition matrix element is determined by the electronic structures including the dipole moments and the excitation energies. So the lowest excitation energy of Zn1e is one of the probable reasons for the strongest TPA intensity among the zinc complexes. Comparing with corresponding free ligands, λtp has a large red-shift and σ is increased greatly for Zn1a, Zn1d, and Zn1e. These results are in good agreement with experimental observations.[22] The arrangement where the metal ion binds to the acceptor site should yield an increased cross section.[19,22]

In the case of probe 2, the first excited state has the largest TPA intensity for all the free ligands and zinc complexes. For 2b, 2c, and 2f, the second excited state also has considerable intensity. The largest cross sections of 2a–2f are all located at about 820 nm with different cross sections. The maximum and the minimum cross sections are 524 GM and 586 GM for 2c and 2e respectively. Both 2b and 2c have two absorption maxima at about 820 nm and 700 nm and the absorption intensities at 700 nm cannot be neglected. As for probe 1, if all the isomers 2a–2f are considered together, the obtained TPA spectrum could have several separated absorption maxima. One can also see that the TPA wavelengths have large red-shifts and the intensities increase greatly due to the stronger electron donor with respect to probe 1.

Table 4.

Value of two-photon absorption wavelength λtp and cross section σ of three lowest excited states of probe 3 in water solvent.

.

Upon Zn2+ binding, each of all the zinc complexes has a strong TPA peak in a longer wavelength range with enhanced cross section in comparison with each of the corresponding free ligands. It shows that the TPA wavelength and cross section are quite different among the zinc complexes. Zn2b has the smallest cross section 596 GM at 838 nm, while Zn2f has the largest value 761 GM at 865 nm. The cross section increases by 165 GM. Like the zinc complexes of probe 1, the TPA intensity is in the order Zn2f>Zn2e>Zn2d>Zn2a>Zn2c>Zn2b. When the zinc ion connects with the nitrogen atoms from the picolyl unit in the middle of the ligand, Zn2e, Zn2d, and Zn2a give rise to the large cross sections in the long wavelength region. Although Zn2f has the largest cross section, the TPA wavelength is located at 865 nm, which is much shorter than those of Zn2e, Zn2d, and Zn2a. According to our calculations, it can be concluded that the coordination mode plays an important role in TPA intensity and wavelength.

The TPA properties of probe 3 are also calculated and the results are shown in Table 4. The data are similar to those of probe 1. Most of the TPA cross sections decrease a little with respect to the corresponding structures of probe 1 due to the weak donor unit.

3.4. Intra-molecular charge transfer

It is well known that TPA properties highly depend on the ICT process. The charge will be redistributed when the molecule is excited from the ground state to the excited state. In order to further explore the effects of the coordination mode and obtain a better understanding of the ICT process, we perform the natural bond orbital charge analyses for Zn1a, Zn1b, Zn1d, and Zn1e in their ground states and in the first excited states. To analyze them clearly, these molecules are divided into three parts as shown in Fig. 4. The net charges of parts A, B, and C are specifically calculated and all the calculated results are listed in Table 5.

Fig. 4. Divided parts of the molecules Zn1a, Zn1b, Zn1d, and Zn1e.
Table 5.

Net charges (in units of e) for divided parts of the molecules in the ground states and in the first excited states.

.

QA0 and QA1 denote the net charge of part A in the ground state and that in the first excited state, respectively. ΔQA represents the net charge difference between the ground state and the first excited state in part A. Like this definition, ΔQB and ΔQC denote the net charge differences between the ground state and the first excited state for part B and part C, respectively. For Zn1a, it is found that the net charge of part A in the first excited state QA1 is −0.102e, which is more electropositive than the value in the ground state QA0, −0.186e. The net charge of part B in the first excited state (QB1, 0.544e) is also more electropositive than in the ground state (QB0, 0.224e). In the case of part C, the net charge in the first excited state (QC1, −0.441e) is more electronegative than in the ground state (QC0, −0.037e). This indicates that part A, together with part B, is the donor, and part C is the acceptor of the molecule. Other structures are similar to those of Zn1a. That is, part A and part B both act as a donor, and part C serves as an acceptor. The net charge difference of the donor should be the sum of ΔQA and ΔQB. For Zn1a, the net charge difference of the donor is 0.404e, which is very close to the value of Zn1d, 0.409e. Zn1b has the smallest value, 0.299e and Zn1e has the biggest one, 0.439e. This demonstrates that the capacity of ICT is in the order Zn1e>Zn1d>Zn1a>Zn1b, which is consistent with the order of TPA intensity.

4. Conclusions

Effects of isomerism and coordination mode on the linear and nonlinear optical properties of three TP probes for zinc ions are theoretically studied by quantum chemical calculations. Six isomers and six corresponding zinc complexes of each probe are optimized and their OPA and TPA properties are calculated at the DFT level. Also, the ICT properties for various coordination conformations are specified. It is found that the isomerism has some effects on TPA properties of free ligands and cannot be neglected. The TPA wavelength and cross section are quite different among the zinc complexes, which demonstrates the significant coordination mode effects on TPA. It is interesting to find the coordination modes where the zinc ion connects with the picolyl unit in the middle of the ligand always give rise to the large cross sections in a long wavelength region. The natural bond orbital charge analysis shows that this coordination conformation has the strongest ICT capacity. Our results provide useful information for understanding the TPA properties of metal ion probes.

Reference
[1] Denk W Strickler J H Webb W W 1990 Science 248 73
[2] Yao S Belfield K D 2012 Eur. J. Org. Chem. 2012 3199
[3] Kim D Ryu H G Ahn K H 2014 Org. Biomol. Chem. 12 4550
[4] Kim H M Cho B R 2015 Chem. Rev. 115 5014
[5] Kim H M Cho B R 2011 Chem. Asian J. 6 58
[6] Sarkar A R Kang D E Kim H M Cho B R 2014 Inorg. Chem. 53 1794
[7] Sumalekshmy S Fahrni C J 2011 Chem. Mater. 23 483
[8] Kim D Singha S Wang T Seo E Lee J H Lee S J Kim K H Ahn K H 2012 Chem. Commun. 48 10243
[9] Kim H M An M J Hong J H Jeong B H Kwon O Hyon J Y Hong S C Lee K J Cho B R 2008 Angew. Chem. Int. Ed. 47 2231
[10] Kim H J Heo C H Kim H M 2013 J. Am. Chem. Soc. 135 17969
[11] Das S K Lim C S Yang S Y Han J H Cho B R 2012 Chem. Commun. 48 8395
[12] Masanta G Heo C H Lim C S Bae S K Cho B R Kim H M 2012 Chem. Commun. 48 3518
[13] Zhang Y Wang J Jia P Yu X Liu H Liu X Zhao N Huang B 2010 Org. Biomol. Chem. 8 4582
[14] Bush A I Pettingell W H Multhaup G Paradis M D Vonsattel J P Gusella J F Beyreuther K Masters C L Tanzi R E 1994 Science 265 1464
[15] Cuajungco M P Lees G J 1997 Neurobiol. Dis. 4 137
[16] Bush A I Tanzi R E 2002 Proc. Natl. Acad. Sci. USA 99 7317
[17] Wang D Guo J F Ren A M Huang S Zhang L Feng J K 2014 J. Phys. Chem. 118 10101
[18] Bednarska J Zaleśny R Murugan N A Bartkowiak W Ågren H Odelius M 2016 J. Phys. Chem. 120 9067
[19] Zhu M Y Zhao K Song J Wang C K 2018 Chin. Phys. 27 023302
[20] Zhang Y J Zhang Q Y Ding H J Song X N Wang C K 2015 Chin. Phys. 24 023301
[21] Zhu M Y Song J Zhao K 2017 J. At. Mol. Sci. 8 152
[22] Sumalekshmy S Henary M M Siegel N Lawson P V Wu Y Schmidt K Brédas J L Perry J W Fahrni C J 2007 J. Am. Chem. Soc. 129 11888
[23] Arnbjerg J Jiménez-Banzo A Paterson M J Nonell S Borrell J I Christiansen O Ogilby P R 2007 J. Am. Chem. Soc. 129 5188
[24] Guillaume M Ruud K Rizzo A Monti S Lin Z Xu X 2010 J. Phys. Chem. 114 6500
[25] Zhao K Liu P W Wang C K Luo Y 2010 J. Phys. Chem. 114 10814
[26] Wang F Q Zhao K Zhu M Y Wang C K 2016 J. Phys. Chem. 120 9708
[27] Jia H H Zhao K Wu X L 2014 Chem. Phys. Lett. 612 151
[28] Zhao K Han G C Zhang L L Jia H H 2014 Chin. J. Chem. Phys. 27 75
[29] Wu X L Zhao K Jia H H Wang F Q 2015 Acta Phys. Sin. 64 233301 in Chinese
[30] Luo Y Norman P Macak P Ågren H 2000 J. Phys. Chem. 104 4718
[31] Olsen J Jørgensen P 1985 J. Chem. Phys. 82 3235
[32] Monson P R McClain W M 1970 J. Chem. Phys. 53 29
[33] GAUSSIAN 09 References in http://www.gaussian.com/
[34] DALTON 2013 A molecular electronic structure program, Release Dalton2013.0 see http://daltonprogram.org/
[35] Xue L Fang Z Li G Wang H Jiang H 2011 Sens. Actuator B-Chem. 156 410