Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter October 20, 2016

Engineering Optical Activity of Semiconductor Nanocrystals via Ion Doping

  • Nikita V. Tepliakov , Anvar S. Baimuratov , Yurii K. Gun’ko , Alexander V. Baranov , Anatoly V. Fedorov and Ivan D. Rukhlenko EMAIL logo
From the journal Nanophotonics

Abstract

Controlling the strength of enantioselective interaction of chiral inorganic nanoparticles with circularly polarized light is an intrinsically interesting subject of contemporary nanophotonics. This interaction is relatively weak, because the chirality scale of nanoparticles is much smaller than the optical wavelength. Here we theoretically demonstrate that ion doping provides a powerful tool of engineering and enhances optical activity of semiconductor nanocrystals. We show that by carefully positioning ionic impurities inside the nanocrystals, one can maximize the rotatory strengths of intraband optical transitions, and make them 100 times larger than the typical rotatory strengths of small chiral molecules.

1 Introduction

The life is known to be essentially chiral, which is why many organic drugs exist in two enantiomeric forms [1, 2]. As these forms often have profoundly different impacts on human body, drug-regulating authorities require them to be separated and tested individually on toxicity and therapeutic efficacy [35]. The detection of enantiomeric forms of chiral molecules is only possible through their interaction with other chiral objects [68]. Semiconductor nanocrystals, which are often intrinsically chiral because of the unavoidable surface and bulk defects, are objects of particular promise owing to their distinctive photoluminescence properties [916]. It has been recently demonstrated that CdSe/ZnS quantum dots and quantum rods can acquire chirality from screw dislocations naturally developing during their growth [1720]. It has been also shown that when semiconductor nanocrystals of one enantiomeric form are mixed with a chiral drug, they can exhibit strong affinity to its particular enantiomer [21]. The detection and quantification of this enantiomer are then possible from the changes in the absorption or emission spectrum of the nanocrystals, which occur due to the energy and charge transfer processes that take place upon binding of the nanocrystals to the targeted enantiomer.

The major challenge of the described detection approach is the fabrication of enantiopure semiconductor nanocrystals. The existing fabrication methods are essentially symmetric, producing mixtures of left-handed, righthanded, and achiral species [17, 18]. The detection of drug enantiomers with such mixtures is still feasible through the analysis of circular dichroism (CD) spectra, comprising the lines of both nanocrystals and drug molecules [22]. The sensitivity of the analysis can be enhanced by strengthening the optical activity of the nanocrystals via their doping with impurity ions such as Mn2+ or Er3+. In this paper, we propose and develop a method of engineering and enhancing the optical activity of semiconductor nanocrystals with specifically positioned ionic impurities.We show that there are certain optimal impurity positions that maximize the rotatory strengths of optical transitions inside the nanocrystals, making them two orders of magnitude larger than the typical rotatory strengths of small chiral molecules. It is also found that the optimally doped quantum dots exhibit an almost total dissymmetry of their absorption upon magnetic dipole transitions.

2 Theoretical model

For a semiconductor nanocrystal to be chiral and optically active, it must lack a plane or center of symmetry. The minimal number of identical impurities required to break the mirror symmetry of a nanocrystal is the largest for a nanosphere. It is easy to see that regardless of their positions, neither one nor two impurities are sufficient to make a nanosphere chiral, because there is always a rotation axis of infinite order associated with one impurity and a mirror plane associated with two. A nanosphere may acquire chirality, provided it is doped with a minimum of three impurity ions, which form a plane that does not pass through the nanosphere’s center. Similarly, a minimum of two impurities are needed to break the mirror symmetry of a nanocylinder, and one to bring chirality to a nanocuboid.

The optical activity of doped semiconductor nanocrystals results from the perturbation of their electronic subsystems by the electric field of the impurities. Figure 1 illustrates this perturbation by the example of two electronic states and one impurity ion inside a nanocuboid. One can see that by breaking the mirror symmetry of the nanocuboid, the impurity makes the electronicwave functions essentially chiral. Thus the induced chirality shows up in the nanocrystal’s CD spectrum

Δεa(ω)=8πNωcnρba(ω)Rba,(1)
Figure 1 Wave functions of electronic states (112) and (224) inside a 3.4 × 3.8 × 4.2 nm3 nanocuboid (a) without and (b) with an impurity ion located at point x = 0.68 nm, y = 0.76 nm, and z = 0.84 nm (shown by dot). For visualization purpose, an ion of charge +10e is used. The wave functions are normalized by the volume of the nanocuboid.
Figure 1

Wave functions of electronic states (112) and (224) inside a 3.4 × 3.8 × 4.2 nm3 nanocuboid (a) without and (b) with an impurity ion located at point x = 0.68 nm, y = 0.76 nm, and z = 0.84 nm (shown by dot). For visualization purpose, an ion of charge +10e is used. The wave functions are normalized by the volume of the nanocuboid.

where N is the concentration of the nanocrystals, c is the speed of light in vacuum, ρab(w) is the spectral lineshape of transition ab, and the rotatory transition strength is defined as [23]

Rba=Im(pabmba),(2)

where p=er,m=iεe/(2m*c)r×,ε is the high-frequency permittivity, and m* and -e are the effective mass and charge of an electron, respectively Note that the so defined rotatory strength does not depend on the polarization of the excitation light and describes ensembles with random nanocrystal orientations such as colloidal quantum dots.

In order to describe the quantum states of electrons confined by a doped semiconductor nanocrystal, we apply a one-band model of bulk semiconductor and ignore the exciton effects by assuming that the characteristic size of the nanocrystal, l, is much smaller than the exciton Bohr radius rBex=ε02/(mexe2), where £0 is the static permittivity and mex is the effective mass of an exciton. We also assume that the nanocrystal is made of a cubic semiconductor with an isotropic material response, characterized by scalar permittivities, and treat the impurities as point elementary charges. If the impurities are located at points Rj, then their impact on the confined electrons is described by an electric potential

V(r)=je2ε0|r-Rj|'(3)

which can be considered as a small perturbation of the nanocrystal. The envelope wave functions and energy states of the confined electrons are characterized by a set of three quantum numbers n and obey the Schrodinger equation

Enψn=(H+V)ψn(4)

in which the Hamiltonian of the impurity-free nanocrystal, H, is achiral.

We next take the solution of Eq. (4) in the form of the standard perturbation expansion En = εn + Vnn and Ψn = ψn + Snn'ψn', where ψn and εn are the wave functions and energy spectrum of the unperturbed Hamiltonian H,Vnn=ψn*Vψndr and we have introduced a summation operator Snn=n'Vnn/εnεn. In writing the expansion in the form shown, we have assumed that the states of the unperturbed Hamiltonian are non-degenerate. It can be shown that this expansion is accurate for a nanocrystal with Z impurity ions, provided that the size of the nanocrystal lrB(3π)2/(2Z), where rB = ε02/(m*e2) is the conductivity electron Bohr radius

Let us focus on studying the optical activity of a semiconductor nanocuboid lx × ly × lz, whose optical activity can be achieved with a single impurity ion. By assuming that the nanocrystal is located symmetrically with respect to the three Cartesian planes and its surface is impenetrable for the confined electrons, we take ψn(r)v2/lvHv(kvv) and εn=2(kx2+ky2+k22)/(2m*), where n = (nx, ny, nz), Hv(ξ) = cos ξ for odd nv, Hv(ξ) = sin ξ for even nv, and kv = knv = πnv/lv. With these wave functions, the i-th components of matrix elements pnn" and mn"n are found to be given by

ψnpiψndr=eli(Bniniδniniδnknk+snnBniniδniniδnknk+snnBniniδniniδnknk(5a)

and

ψnmiψndr=iεe2mcεijk(δniniGnknknknk+SnnδniniGnknknknk+SnnGnknknknk),(5b)

where ijk, δij and εijk are the Kronecker's delta and the Levi-Civita symbol, and

Bnini=8niniπ2(n12n2i)2Re ini+ni+1,(6)
Gnknknknk=ljlk2BnjnjBnknk(knj2knj2knk2+knk2),(7)

The selection rules for intraband transitions are determined by Kronecker deltas, coefficients Bn1n1 (which are nonzero only when ni and n1 are of different parities), and coefficients Gnjnknknk. It is readily seen that Rba = 0 if the im-purities are absent, because of the different selection rules for the electric and magnetic moment operators.

The obtained result leads to a simple size dependence of the rotatory strength. Suppose that the lengths of the nanocrystal edges and the coordinates of the impurities are increased by a factor of f. Then εnf-2, Vf-1 and Snn'f. As the matrix element of the magnetic dipole moment does not change with f while the electric dipole moment scales as ∝ f, Eqs. (2) and (5) give Rbaf2. One can see that the size dependence of the rotatory strength comes from three sources: scaling of the confinement energies, scaling of the impurity potential, and scaling of the electric dipole moment. As the nanocrystal grows bigger, the Coulomb interaction between electrons and holes more and more reduces the confinement energies, resulting in their weaker size dependence εnf-q, where 1 < q < 2. Hence, the rotatory strength in the weak confinement regime is larger than that in the strong confinement regime but has a weaker size dependence, Rbafq.

Equations (2) and (5) show that there are six types of dipole-allowed transitions, two per each Cartesian axis. In particular, the confined motion along the x-axis leads to electric dipole transitions, which change the parity of state ψ ψx and preserve states ψy and ψz, and to magnetic dipole transitions, which preserve state ψx and change the parities of states ψy and ψz. The electric dipole mo-ment is much larger than the magnetic dipole moment for the transitions of the first kind, |pab | ≫ |mba|, and the opposite relation holds for the transitions of the second kind, |mha| ≫ |Pab|.

The rotatory strength of transition (nx,ny,nz)(nx,ny,nz) induced by the x components of the electric and magnetic dipole moments is determined by the expressions

Bnxnx(Snxnynz;nxnynzSnxnynz;nxnynz)Gnynynznz,(8a)
Gnynynznz(Snxnynz;nxnynzSnxnynz;nxnynz)Bnxnx.(8b)

The rotatory strength is seen to be contributed only by the sums involving the matrix elements of the perturbation-induced transitions between the states with different parities of the respective quantum numbers. For a nanocrys-tal doped with one impurity ion, these matrix elements are antisymmetric with respect to each ion's coordinate, making rotatory strength vanish in case the ion is located in one of the three Cartesian planes. This conclusion also holds for other components of the dipole moments, the rotatory strengths of which can be obtained from Eq. (8) by a cyclic change of subscripts x, y, and z. As the reflection of the impurity in any of the coordinate planes changes the sign of the rotatory strength, knowing Rba for all impurity locations in the first octant of the nanocrystal allows us to find Rba in the entire nanocrystal.

3 Results and discussion

Engineering optical activity of doped nanocrystals requires knowing the optimal positions ofimpurityions that maximize the activity. The optimal positions critically depend on the transition upon which the activity is observed, and can be estimated by analyzing the dominant terms in Eq. (8). These terms correspond to the perturbation-induced transitions between the unperturbed initial state ψx, ny, nz and the unperturbed final state ψnx,ny,nz and the closest to them unperturbed states of the nanocrystal. To illustrate this point, consider six transitions from the ground state to states Ψ211, Ψ121, Ψ112, Ψ122, Ψ212, and Ψ221. For all of them, ψ222 is the closest state coupled by perturbation V to state ψ111. The perturbation-induced transitions between these states give the first dominant contribution to the six rotatory strengths. If the three nanocrystal edges do not differ too much, then the second major contribution to R211;111, R121;111, ⋯ comes from transitions ψ211ψ122,ψ121ψ212, Thema-trix elements of these transitions are all alike, V111;222 = V211;122 = ..., resulting in the same dependence of the six dominant terms on the impurity position. The locations and number of absolute maxima of rotatory strengths R211;111, R121;111, ... are therefore, determined by a single matrix element V111;222. As the wave function product vcos(πv/lv)sin(2πv/lv) has one maximum in the first octant, there is also a single absolute maximum of the six rotatory strengths.

The isosurfaces of Rba shown in Fig. 2 illustrate the dependence of rotatory strength on the impurity position inside the first octants of nanorod, quantum dot, and nanoplatelet of approximately equal volumes. In agreement with the above discussion, the rotatory strength upon the electric dipole transition Ψ111Ψ121 peaks at one point, clearly seen in panels (a)-(c). The optimal impurity positions inside the nanorod, quantum dot, and nanoplatelet are (0.50, 0.44, 0.40), (0.46, 0.44, 0.42), and (0.50, 0.42, 0.40), respectively. According to panel (d), rotatory strength R141;111 has two maximums located at points (0.46, 0.22, 0.36) and (0.46, 0.72, 0.36) along the y-axis; they come from two maximums of the wave function product cos(πuy|2)sin(2πuy). Similarly, there are two and three maximums of R124;111 and R116;111 along the z-direction, as can be seen from panels (e) and (f).

Figure 2 Isosurfaces of rotatory strength Rba (in the units of 10-39 erg × cm3) in the first octant of 2 × 3 × 9 nm3 nanorod, 3.4 × 3.8 × 4.2 nm3 quantum dot, and 1.5 × 5 × 6 nm3 nanoplatelet. Transitions occur from the ground state (111) to state (121) in panels (a)-(c) and to states (141), (124), and (116) in panels (d), (e), and (f), respectively. All the plots are in normalized coordinates uv = 2v|lv (v = x, y, z). The modeling material is InP and the simulation parameters are: ε0 = 12.35, ε = 9.61, m* = 0.08 m0 (m0 is the free-electron mass), and rB = 8.12 nm.
Figure 2

Isosurfaces of rotatory strength Rba (in the units of 10-39 erg × cm3) in the first octant of 2 × 3 × 9 nm3 nanorod, 3.4 × 3.8 × 4.2 nm3 quantum dot, and 1.5 × 5 × 6 nm3 nanoplatelet. Transitions occur from the ground state (111) to state (121) in panels (a)-(c) and to states (141), (124), and (116) in panels (d), (e), and (f), respectively. All the plots are in normalized coordinates uv = 2v|lv (v = x, y, z). The modeling material is InP and the simulation parameters are: ε0 = 12.35, ε = 9.61, m* = 0.08 m0 (m0 is the free-electron mass), and rB = 8.12 nm.

One can see that the optical activity of semiconductor nanocrystals doped with one ion of charge +e can be 10 times larger than the rotatory strengths of chiral molecules, which typically range from 10-39 to 10-38 erg × cm3 [24-26]. Using several dopants of bigger charges, optimally positioned in the third, sixths, and eighths octants, one can increase the rotatory strengths of optical transitions by another factor of 10. It should also be noted that the rotatory strength is maximal for the lowest-energy intraband transitions and steeply decreases with transition energy becouse of the decrease in the probabilities of the electric and magnetic dipole transitions, governed by the coefficients Bnini and Gnjnjnknk

Figure 3 CD spectra of (a) 2 × 3 × 9 nm3 nanorod, (b) 3.4 × 3.8 × 4.2 nm3 quantum dot, and (c) 1.5 × 5 × 6 nm3 nanoplatelet with different positions of an impurity shown by the color dots in the insets. The insets represent cross sections of the nanocrystals at x = lx/4. All the spectral lines are approximated by Lorentzians with a FWHM of 0.1 eV [27]. The material parameters are the same as in Fig. 2.
Figure 3

CD spectra of (a) 2 × 3 × 9 nm3 nanorod, (b) 3.4 × 3.8 × 4.2 nm3 quantum dot, and (c) 1.5 × 5 × 6 nm3 nanoplatelet with different positions of an impurity shown by the color dots in the insets. The insets represent cross sections of the nanocrystals at x = lx/4. All the spectral lines are approximated by Lorentzians with a FWHM of 0.1 eV [27]. The material parameters are the same as in Fig. 2.

Despite the fact that the rates of electric dipole and magnetic dipole transitions differ by a factor of 10,000 [28], the respective rotatory strengths are of the same order of magnitude. The large difference in the transition rates shows up in dissymmetry factor g, which is the doubled ratio of CD to the sum of absorption intensities for left- and right-handed circularly polarized light. This feature is illustrated in Table 1. The large absorption upon the electric dipole transition is seen to result in g-factors in the range from 10-4 to 10-3, which is typical for small chiral molecules [29]. As the magnetic dipole absorption is about 104 weaker than the electric dipole one, the g-factors of the magnetic dipole transitions can be of the order of unity, indicating a strong chiroptical response. Note that the g-factors in Table 1 are not maximal, as they are calculated for the impurity position providing maximum to the rotatory strength.

Table 1

Rotatory strengths (in the units of erg × cm3) and dissymmetry factors of (a) electric dipole transition (111) → (121) and (b) magnetic dipole transition (111) → (122) inside nanorod, quantum dot, and nanoplatelet with a single impurity located at optimal position u, which corresponds to the maximum of rotatory strength. All material parameters are the same as in Fig. 2.

NanorodQuantum dotNanoplatelet
u(a)(0.50,0.44,0.40)(0.46,0.44,0.42)(0.50,0.42,0.40)
R(a)-136 × 10-39-60 × 10-39-19 × 10-39
g(a)8.1 × 10-42.2 × 10-40.4 × 10-4
u(b)(0.50,0.44,0.40)(0.46,0.44,0.42)(0.50,0.42,0.40)
R(b)156 × 10-39-59 × 10-397 × 10-39
g(b)-0.1861.150-0.324

The dependence of the CD spectra of doped nanocrys-tals on the impurity position is illustrated in Fig. 3. The impurity is assumed to be in plane x = lx/4, where its positions are shown by the color dots in the insets. The optical CD spectrum of the nanorod in Fig. 3(a) has two strong peaks, corresponding to the transitions in Table 1, and a much weaker peak of transition Ψ111Ψ124. As the impurity is moved from left to right, the weaker peak changes its sign, while nearly disappearing when the dominant peaks have maximum intensity. The CD spectrum of the quantum dot in Fig. 3(b) consists of six equally pro-nounced dipole transitions, whose rotatory strengths were shown to exhibit almost the same dependence on the impurity position. All of the peaks are the strongest where the impurity is located about the center of the first octant. The optical CD spectrum of the nanoplatelet shown in Fig. 3(c) includes four transitions, whose intensities depend on the impurity position as it is predicted by Fig. 2(d) and (e). There are also three low-energy peaks in the spectrum, which vary with position similar to the states of the quantum dot.

4 Conclusion

We have shown that one can control and maximize the optical activity of semiconductor nanocrystals by introducing impurity ions inside them. The proposed approach allows one to achieve rotatory strengths upon intraband transitions that are 100 times larger than the typical rotatory strengths of small chiral molecules. It also enables re-alization of an almost complete dissymmetry of magnetic dipole absorption, opening a new chapter in the studies of light-matter interaction on the nanoscale.

Acknowledgement

This work was funded by Grant 14.B25.31.0002 and Government Assignment No.3.17.2014/K of the Ministry of Education and Science of the Russian Federation. A.S.B. acknowledges the Dynasty Foundation Support Program for Physicists and the scholarship of the President of the Russian Federation for young scientists. We also acknowledge the financial support from Science Foundation Ireland (Grant SFI12/IA/1300).

References

[1] L. A. Nguyen, H. He, and C. Pham-Huy, “Chiral drugs: An overview,” Inter. J. Biomed. Sci. 2, 85-100 (2006).Search in Google Scholar

[2] R. N. Compton and R. M. Pagni, “The chirality of biomolecules,” Adv. Atom. Mol. Opt. Phys. 48, 219-261 (2002).10.1016/S1049-250X(02)80010-6Search in Google Scholar

[3] W. H. Brooks, W. C. Guida, and K. G. Daniel, “The significance of chirality in drug design and development,” Curr. Top. Med. Chem. 11, 760-770 (2011).10.2174/156802611795165098Search in Google Scholar

[4] B. Li and D. T. Haynie, “Chiral drug separation,” Encycl. Chem. Process. 1, 449-458 (2006).Search in Google Scholar

[5] H. Caner, E. Groner, L. Levy, and I. Agranat, “Trends in the development of chiral drugs,” Drug Discov. Today 9, 105-110 (2004).10.1016/S1359-6446(03)02904-0Search in Google Scholar

[6] G. Gübitz and M. G. Schmid, “Chiral separation principles in chromatographic and electromigration techniques,” Mol. Biotech. 32, 159-179 (2006).10.1385/MB:32:2:159Search in Google Scholar

[7] C.J. Welch, Chiral Chromatography in Support of Pharmaceutical Process Research, Preparative Enantioselective Chromotog-raphy (ed. G. B. Cox) (Blackwell, London, UK, 2005).Search in Google Scholar

[8] S. F. Mason, Molecular Optical Activity and the Chiral Discriminations (Cambridge Univ. Press, New York, 1982).Search in Google Scholar

[9] A. Hazarika, A. Layek, S. De, A. Nag, S. Debnath, P. Mahadevan, A. Chowdhury, and D. D. Sarma, “Ultranarrow and widely tunable Mn2+-induced photoluminescence from single Mn-doped nanocrystals of ZnS-CdS alloys,” Phys. Rev. Lett. 110, 267401 (2013).Search in Google Scholar

[10] E. V. Ushakova, A. P. Litvin, P. S. Parfenov, A. V. Fedorov, M. Arte-myev, A. V. Prudnikau, I. D. Rukhlenko, and A. V. Baranov, “Anomalous size-dependent decay of low-energy luminescence from PbS quantum dots in colloidal solution,” ACS Nano 6, 8913-8921 (2012).10.1021/nn3029106Search in Google Scholar

[11] V. K. Turkov, S. Y. Kruchinin, and A. V. Fedorov, “Intraband optical transitions in semiconductor quantum dots: Radiative electronic-excitation lifetime,” Opt. Spectrosc. 110, 740-747 (2011).10.1134/S0030400X11050183Search in Google Scholar

[12] W. M. Witzel, A. Shabaev, C. S. Hellberg, V. L. Jacobs, and A. L. Efros, “Quantum simulation of multiple-exciton generation in a nanocrystal by a single photon,” Phys. Rev. Lett. 105, 137401 (2010).10.1103/PhysRevLett.105.137401Search in Google Scholar

[13] R. D. Schaller, S. A. Crooker, D. A. Bussian, J. M. Pietryga, J. Joo, and V. I. Klimov, “Revealing the exciton fine structure of PbSe nanocrystal quantum dots using optical spectroscopy in high magnetic fields,” Phys. Rev. Lett. 105, 067403 (2010).10.1103/PhysRevLett.105.067403Search in Google Scholar

[14] A. V. Fedorov, A. V. Baranov, I. D. Rukhlenko, T. S. Perova, and K. Berwick, “Quantum dot energy relaxation mediated by plas-mon emission in doped covalent semiconductor heterostruc-tures,” Phys. Rev. B 76, 045332 (2007).10.1103/PhysRevB.76.045332Search in Google Scholar

[15] I. D. Rukhlenko and A. V. Fedorov, “Propagation of electric fields induced by optical phonons in semiconductor heterostructures,” Opt. Spectrosc. 100, 238-244 (2006).10.1134/S0030400X06020135Search in Google Scholar

[16] I. D. Rukhlenko and A. V. Fedorov, “Penetration of electric fields induced by surface phonon modes into the layers of a semicon-ductor heterostructure,” Opt. Spectrosc. 101, 253-264 (2006).10.1134/S0030400X06080133Search in Google Scholar

[17] M. P. Moloney, J. Govan, A. Loudon, M. Mukhina, and Y. K. Gun'ko, “Preparation of chiral quantum dots,” Nature Protoc. 10, 558-573 (2015).10.1038/nprot.2015.028Search in Google Scholar PubMed

[18] M. V. Mukhina, V. G. Maslov, A. V. Baranov, A. V. Fedorov, A. O. Orlova, F. Purcell-Milton, J. Govan, and Y. K. Gun'ko, “Intrinsic chirality of CdSe/ZnS quantum dots and quantum rods,” Nano Lett. 15, 2844-2851 (2015).10.1021/nl504439wSearch in Google Scholar PubMed

[19] A. S. Baimuratov, I. D. Rukhlenko, R. E. Noskov, P. Ginzburg, Y. K. Gun'ko, A. V. Baranov, and A. V. Fedorov, “Giant optical activity of quantum dots, rods, and disks with screw dislocations,” Sci. Rep. 5, 14712 (2015).10.1038/srep14712Search in Google Scholar PubMed PubMed Central

[20] A. S. Baimuratov, I. D. Rukhlenko, Y. K. Gun'ko, A. V. Baranov, and A. V. Fedorov, “Dislocation-induced chirality of semiconductor nanocrystals,” Nano Lett. 15, 1710-1715 (2015).10.1021/nl504369xSearch in Google Scholar PubMed

[21] S. D. Elliot, M. P. Moloney, and Y. K. Gun'ko, “Chiral shells and achiral cores in CdS quantum dots,” Nano Lett. 8, 2452-2457 (2008).10.1021/nl801453gSearch in Google Scholar PubMed

[22] N. Berova, Comprehensive Chiroptical Spectroscopy, vol. 1 (John Wiley & Sons, 2012).10.1002/9781118120392Search in Google Scholar

[23] L. Rosenfeld, “Quantenmechanische theorie der natürlichen optischen aktivität von flüssigkeiten und gasen,” Z. Phys. 52, 161174 (1929).10.1007/BF01342393Search in Google Scholar

[24] R. R. Judkins and D. J. Royer, “Optical rotatory strength of tris-bidentate cobalt(II1) complexes,” Inorganic Chem. 13, 945-950 (1974).10.1021/ic50134a037Search in Google Scholar

[25] T. S. King, P. M. Bayley, and F. C. Yong, “Optical rotatory dispersion and circular dichroism of cytochrome oxidase,” Eur. J. Biochem. 20, 103-110 (1971).10.1111/j.1432-1033.1971.tb01367.xSearch in Google Scholar PubMed

[26] P. M. Bayley, E. B. Nielsen, and J. A. Shellman, “The rotatory properties of molecules containing two peptide groups: The-ory,” J. Phys. Chem. 73, 228-243 (1969).10.1021/j100721a038Search in Google Scholar PubMed

[27] O. I. Micic, J. Sprague, Z. Lu, and A. J. Nozik, “Highly efficient band-edge emission from In P quantum dots,” Appl. Phys. Lett. 68, 3150-3152 (1996).10.1063/1.115807Search in Google Scholar

[28] H. Giessen and R. Vogelgesang, “Glimpsing the weak magnetic field of light,” Science 326, 529 (2009).10.1126/science.1181552Search in Google Scholar PubMed

[29] M. Wakabayashi, S. Yokojima, T. Fukaminato, K. Shiino, M. Irie, and S. Nakamura, “Anisotropic dissymmetry factor, g: Theoretical investigation on single molecule chiroptical spectroscopy,” J. Phys. Chem. A 118, 5046-5057 (2014).10.1021/jp409559tSearch in Google Scholar PubMed

Received: 2016-1-22
Accepted: 2016-2-13
Published Online: 2016-10-20
Published in Print: 2016-9-1

© 2016 Nikita V. Tepliakov et al., published by De Gruyter Open

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 3.6.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2016-0034/html
Scroll to top button