Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Plasmon-hybridization-induced optical torque between twisted metal nanorods

Open Access Open Access

Abstract

We present a numerical study of optical torque between two twisted metal nanorods due to the angular momentum of the electromagnetic field emerging from their plasmonic coupling. Our results indicate that the interaction optical torque on the nanorods can be strongly enhanced by their plasmon coupling, which highly depends on not only the gap size but also the twisted angle between the nanorods. The behaviors of the optical torque are different between two plasmon coupling modes: hybridized bonding and anti-bonding modes with different resonances. The rotations of the twisted nanorods with the bonding and anti-bonding mode excitations lead to mutually parallel and perpendicular alignments, respectively. At an incident intensity of 10 mW/μm2, the rotational potential depths are more than 30 times as large as the Brownian motion energy, enabling the optical alignments with angle fluctuations less than ∼±10°. Thus, this optical alignment of the nanoparticles with the plasmon coupling allows dynamic control of the plasmonic characteristics and functions.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Noble metal nanoparticles exhibit a strong interaction with light due to the excitation of collective oscillations of their free electrons, which is known as localized surface plasmon resonance (LSPR) [13]. On the resonance, the plasmon oscillation gives rise to an intense near field on the nanoparticle surface and a strong scattering light in the far field. The plasmonic characteristics strongly depend on not only the nanoparticle size and shape but also the configurations of the particles, e.g., their separation, orientation and so on [47]. In particular, when two nanoparticles are close to each other, their plasmon coupling occurs owing to the strong interactions [57], resulting in an appearance of two plasmon modes different from individual constituents. The plasmon resonances and the near-field enhancement effects of these hybridized modes are significantly changed by the nanoparticle configurations [812]. In other words, we can efficiently control the plasmon characteristics through the configurations with the plasmon coupling.

The plasmon coupling between nanoparticles drastically enhances the electromagnetic (EM) field in the narrow gap separating them. Over many decades, extensive investigations on the basis of plasmon coupling have been carried out, promoting the applications of plasmonics, such as surface-enhanced fluorescence [13,14], surface-enhanced Raman scattering [15,16], chemical sensors and biosensors [17,18], and high harmonic generation [19,20]. Furthermore, the plasmon coupling strongly enhances optical forces between two nanoparticles [21,22]. This interaction force provides a possible approach for assembly control of the nanoparticles [2325].

The optical forces are induced by the linear momentum transfer between light and matter, and they have been widely applied to optical manipulations [2632]. Light has not only the linear momentum but also angular momentum [3238]. The angular momentum carried by light can produce an optical torque via scattering or absorption, which has been firstly demonstrated by Beth more than 80 years ago [38]. The optical torque is able to rotate objects in a controlled manner. It has been attracting widespread attention since it improves the mechanical degree of freedom to manipulate the objects, with a variety of applications in atomic and molecular physics, nanotechnology and biology [3944].

Here, we study, for the first time, the direct relation between the optical torque and plasmon hybridization between twisted metal nanorods. Our numerical simulations indicate that the behaviors of the interaction optical torque at hybridized modes are different from that of an isolated nanorod, which depend on not only the gap size but also the twisted angle between the nanorods. This interaction optical torque implements the rotations to mutually perpendicular and parallel alignments of nanorods with the light excitations of different hybridized modes. Thus, the torques due to the plasmon hybridization would play an important role to control the plasmonic characteristics and functions.

2. Methods

Optical torque was studied using the finite element method simulations with the Maxwell stress tensor (MST). Another approach based on the dipole approximation is also presented in Appendix Section A.1. In order to simply excite and observe the plasmon hybridized modes, we chose a dimer of twisted nanorods as shown in Fig. 1. The two nanorods are separated by gap size d, and they are twisted by angle θ. The diameter (D) and length (L) of both nanorods were 40 nm and 120 nm, respectively. The long axis of Rod 1 in Fig. 1 was fixed along the y-axis. Conversely, Rod 2 was rotatable for changing the twisted angle. The dimer was irradiated by a y-polarized plane wave to propagate along the z-axis from the positive to the negative direction. We chose the linearly polarized light without any intrinsic angular momentum, which simplifies the analysis of physical mechanism on the optical torque. The surrounding medium was air. The optical torque on the nanorods was calculated from the conservation law of angular momentum, which can be expressed as [45]

$$\frac{d}{{dt}}\textrm{(}{{\textbf J}_{\textrm{mech}}}\textrm{ + }{{\textbf J}_{\textrm{field}}}\textrm{)} ={-} \int\!\!\!\int_{{S_\textrm{r}}} {\textrm{(}{{\textbf T}_\textrm{M}}} { \times }{\textbf r}\textrm{)} \cdot {{\textbf n}_\textrm{r}}d{S_\textrm{r}},$$
where Jmech is the mechanical angular momentum of the nanorods, Jfield is the angular momentum of the field, Sr is an arbitrary closed surface surrounding the nanorods, and nr is the outward normal unit vector to its surface. ${{\textbf T}_\textrm{M}}$ is the MST, which is defined as
$${{\textbf T}_\textrm{M}} = \varepsilon {\textbf E} \otimes {\textbf E} + \frac{\textrm{1}}{\mu }{\textbf B} \otimes {\textbf B} - \frac{\textrm{1}}{\textrm{2}}\textrm{(}\varepsilon {\textbf E} \times {\textbf E} + \frac{\textrm{1}}{\mu }{\textbf B} \times {\textbf B}\textrm{)}{\textbf I,}$$
where ɛ and μ are the permittivity and permeability of the surrounding medium. ⊗ denotes the operation of dyadic product. I is the unit dyadic. E and B are the electric and magnetic fields, respectively. The right-hand side of Eq. (1) represents the flux of angular momentum that enters the surface Sr. The time derivative of the field angular momentum is zero when it is averaged over one oscillation period. Therefore, the time-averaged mechanical torque on the nanorods can be obtained by
$${\boldsymbol N} = \textrm{ } - \int\!\!\!\int_{{S_\textrm{r}}} {\textrm{(}{{\bar{{\textbf T}}}_\textrm{M}}} { \times }{\textbf r}\textrm{)} \cdot {{\textbf n}_\textrm{r}}d{S_\textrm{r}},$$
where ${\bar{{\textbf T}}_\textrm{M}}$ is the time-averaged MST. We discuss the optical torque only along the z-axis because the torques along the x-and y-axes are negligibly small.

 figure: Fig. 1.

Fig. 1. Schematic illustration of twisted gold nanorods with gap size d and twisted angle θ. The diameter (D) and length (L) of both nanorods are 40 nm and 120 nm, respectively.

Download Full Size | PDF

3. Results and discussion

In the twisted nanorod dimer, the gap size d and twisted angle θ are crucial parameters to decide its configuration. Figure 2(a) shows the plasmon resonance spectra of an isolated nanorod and its dimer at a twisted angle of π/12 with different gap sizes d. As the gap size decreases, the resonance of individual nanorods, i.e., the isolated nanorod, starts to split and two resonance peaks shift to shorter and longer wavelengths due to the plasmon coupling between the nanorods. We also observe similar spectral behavior in the twisted angle dependence for the dimer with a gap size of 1 nm, as shown in Fig. 2(b). As the twisted angle increases and decreases, the separation between two resonance peaks decreases and increases, respectively. At the twisted angle of π/2, the dimer exhibits only one resonance peak even with the gap size of 1 nm, in other words, no plasmon coupling occurs. These results show that the plasmon coupling between twisted nanorods can be controlled by not only the gap size but also the twisted angle.

 figure: Fig. 2.

Fig. 2. Plasmon resonance spectra of an isolated nanorod and its dimer: (a) gap size dependence (θ = π/12) and (b) twisted angle dependence (d = 1 nm). The solid lines, dashed lines, and dotted lines show the extinction, absorption and scattering cross sections, respectively. The vertical dashed lines show the resonance peak wavelength of the isolated nanorod.

Download Full Size | PDF

The coupling between two nanorods has been described as an elegant physical picture which is plasmon hybridization in analogy with molecular orbital theory [812]. Each nanorod can be approximated as a dipole with energy U. The plasmon coupling between two dipole modes generates two new plasmon modes: anti-bonding mode and bonding mode (see Fig. 3). These two hybridized modes are corresponding to a higher energy U+ and a lower energy U respectively, which are produced by splitting from the degenerate energy levels. For the anti-bonding mode, two dipoles are arranged in a parallel manner, i.e., in phase, leading to a constructive superposition of the dipole moment. It is therefore referred to as a bright plasmon mode that shows a large cross section (see Fig. 2). Also, the charges with the same sign at the edges of nanorods result in a higher energy mode due to the charge repulsion. However, for the binding mode, the dipoles are in an anti-parallel alignment that shows opposite charges at the edges of the nanorods, reducing the charge repulsion and leading to a lower energy. In this case, the destructive superposition of the dipole moment produces a dark plasmon mode that exhibits a small cross section (see Fig. 2). In the dimer of twisted nanorods (see Fig. 3), the decrease of the twisted angle, as well as the gap size, would relieve the charge repulsion for bonding mode but inverse for anti-bonding mode, resulting in the LSPR redshift and blueshift, respectively. Additionally, at the twisted angle of π/2, the charge repulsion between two nanorods vanishes away due to the orthogonal dipoles, so that plasmon hybridization does not occur.

 figure: Fig. 3.

Fig. 3. Plasmon hybridization diagram in nanorod dimer with twisted angle θ. The polarization direction of the incident light is along the longitudinal axis of the bottom nanorod.

Download Full Size | PDF

Figure 4 shows the optical torque applied to such twisted nanorods. The behaviors of the optical torque strongly depend on the gap size and the twisted angle between the nanorods. The magnitude of the torque on the whole dimer at the resonance wavelength is similar to that on the isolated nanorod, even in the case of the strong plasmon coupling between the nanorods, i.e., small gap size and twisted angle. However, the optical torque on each nanorod in the dimer can be remarkably enhanced by the plasmon coupling, leading to more than 10 times as large as that on the isolated nanorod. It should be noted that the extinction cross section at the redshifted plasmon resonance (bonding mode) is much smaller than that at the blueshifted plasmon resonance (anti-bonding mode) in Fig. 2, whereas the magnitude of the optical torque at the redshifted resonance is much larger than that at the blueshifted resonance in Figs. 4(b) and 4(d).

 figure: Fig. 4.

Fig. 4. Optical torque acting on (a, c) the whole dimer and (b, d) on each nanorod: (a, b) gap size dependence (θ = π/12) and (c, d) twisted angle dependence (d = 1 nm). The dashed and dotdashed lines show the peak wavelengths at the redshifted resonance. The left- and right- side insets in Fig. 4(d) show the rotation directions of Rod 2 at the blueshifted and redshifted resonances, respectively.

Download Full Size | PDF

Let us discuss the spectral shape and the mechanism of the optical torques on the twisted nanorods. When a linearly polarized field E illuminates an isolated nanorod, the time-averaged optical torque on the induced dipole moment p of the nanorod can be simply expressed as [46,47]

$${{\boldsymbol N}_i} = \frac{\textrm{1}}{\textrm{2}}\textrm{Re[}{\textbf p} \times {{\textbf E}^ \ast }\textrm{],}$$
where ∗ denotes the operation of conjugation. This optical torque is induced by the generation of the angular momentum owing to the interference between scattering light from the dipole moment p and the incident light field E. For longer and shorter wavelengths of the incident light than the plasmon resonance peak in Fig. 2, the nanorod experiences the negative and positive torque, respectively, as shown in Fig. 4(a) (also see Appendix Section A.2), because the phase difference between the dipole moment p and the incident light E changes from 0 to π around the resonance peak. At the resonance peak, the total angular momentum of the interference field is 0, hence, there is no optical torque on the nanorod. In addition, when the incident field is polarized parallel or perpendicular to the long axis of the nanorod, i.e., θ = 0, π/2, the optical torque disappears because of no generation of the angular momentum. In the case of the twisted nanorod dimer, the optical torque in Figs. 4(a) and 4(c) arises from two different mechanisms. One of them is due to the angular momentum in the interference field between the scattering light and the incident light, which is similar to that for the isolated nanorod. Another is due to the different absorption cross section of the left-handed and the right-handed circularly polarized light with positive and negative spin angular momentums, respectively, because of the chirality of the dimer (θ ≠ 0, π/2), which is highly associated with the plasmon coupling in the dimer. On the redshifted resonance, the optical torque induced by the latter is dominant because of the negligible small scattering cross section of the bonding mode (see Fig. 2); therefore, its spectral shape is similar to the absorption cross section. On the blueshifted resonance, the optical torques induced by the former and the latter are mixed, resulting in the complex spectral behavior.

For each nanorod in the dimer, the optical torque is determined by the scattered field from the other nanorod together with the incident light. In Figs. 4(b) and 4(d), the directions of optical torque on the two nanorods are opposite while their magnitudes are close to each other. In other words, each nanorod dominantly experiences an optical torque due to the angular momentum of the EM field arising from the interaction between the two nanorod-plasmons, i.e., interaction optical torque. Therefore, the sign of the optical torque on the nanorods should be decided by the phase difference between the dipole moments of the two nanorods. For the anti-bonding mode (blueshifted resonance), the two dipoles are in phase, producing a negative torque on Rod 1 and a positive torque on Rod 2. Conversely, for the bonding mode (redshifted resonance), the dipoles are out of phase. The optical torques on Rod 1 and Rod 2 are positive and negative, respectively. Moreover, because the bonding plasmon mode has much smaller scattering loss than the anti-bonding mode in Fig. 2, its near field between two nanorods can be more strongly enhanced, as shown in Fig. 5. This results in a much larger magnitude of the interaction optical torque on the redshifted resonance even with its smaller extinction cross section. Additionally, in the case of the two nanorods parallel and perpendicular to each other, i.e., θ = 0, π/2, the EM field does not possess the angular momentum, resulting in no optical torque on the nanorods. Such spectral behavior of the optical torque on each nanorod can be confirmed by another approach based on the dipole approximation (see Appendix Section A.1 for more detail).

 figure: Fig. 5.

Fig. 5. (a) Twisted angle dependence of electric field intensity enhancement at the gap center (d = 1 nm). The dot-dashed lines show the peak wavelengths at the redshifted resonance. (b, c) Distributions of the intensity enhancement at the resonance wavelength of 610 nm and 885 nm, respectively, with the twisted angle of π/12 at the gap center plane.

Download Full Size | PDF

As shown in the inset of Fig. 4(d), the optical torque on Rod 2 for the blueshifted resonance produces anti-clockwise rotations, whereas the torques for the redshifted resonance yield the opposite rotations. As the Rod 1 is fixed along the y-axis, the former and the latter lead to mutually perpendicular and parallel alignments of the nanorods, i.e., θ = π/2 and 0, respectively. For this optical alignment dependent on the incident light wavelength, the torque on the Rod 2 should be large enough to overcome its rotational Brownian motion. The average energy of the Brownian motion associated with each degree of freedom can be expressed as kBT/2 in a system at thermodynamic equilibrium, where kB is the Boltzmann constant and T is the absolute temperature [45]. In this system, the temperature T will increase due to the light absorption of the nanorods which depends on their configurations. To simplify the discussion, we keep T at 300 K. We define the rotational potential energy as

$${U_\textrm{p}}\textrm{(}\theta \textrm{)} ={-} \int_{{\theta _\textrm{0}}}^\theta {Nd\theta } ,$$
where θ0 is π/2 and 0 for the redshifted and blueshifted resonance, respectively.

Figure 6 shows the rotational potentials on Rod 2 for its twisted angle θ at different incident light wavelengths on the redshifted and blueshifted resonances. The Rod 1 is fixed along the y-axis. Supercontinuum white lights with a uniform power over the spectrum from 500 nm to 645 nm (blueshifted resonance) and 645 nm to 1000 nm (redshifted resonance) are assumed to be the incident lights. Their illumination with the same intensity generates the potential wells with similar depths, which show the potential minimums at different twisted angles. To identify the equilibrium positions, we should compare the rotational potential energy with the Brownian motion energy. For an incident intensity of 10 mW/μm2, the potential depths are more than 30 times as large as the Brownian motion energy kBT/2. These potential wells enable the perpendicular and parallel alignments of the nanorods with angle fluctuations of ~±9° and ~±4°, respectively.

 figure: Fig. 6.

Fig. 6. Rotational potentials in units of kBT (T = 300K) for Rod 2 in the dimer (d = 1 nm) as a function of the twisted angle at different incident light wavelengths. Supercontinuum white lights with a uniform power over the spectrum from 500 nm to 645 nm (blueshifted resonance) and 645 nm to 1000 nm (redshifted resonance) are assumed to be the incident lights.

Download Full Size | PDF

To achieve such optical alignments in an experiment, the Rod 1 should be embedded on the surface of a transparent substrate, e.g., silica (see Appendix Section A.3). Furthermore, the sign of optical force, consisting of interaction optical force and radiation pressure, on the Rod 2 should be considered to discuss the possibility of the optical alignments. In the configuration in Fig. 1, the Rod 2 experiences optical force toward the Rod 1 over the wavelength of the incident light (see Appendix Section A.4 for more detail). Thus, the Rod 2 in a solution dropped on the substrate surface can approach the Rod 1. Additionally, we can choose other metals that can further enhance the interaction optical torque to facilitate the optical alignment. For example, silver is another common noble metal to provide LSPRs in the visible spectral range, which is well separated from its interband transitions in comparison with the gold LSPRs [1]. The plasmon resonance of the silver dimer is sharper than that of the gold dimer because of its low ohmic losses and electronic interband transitions. This further enhances the peak magnitude of the interaction optical torque.

4. Conclusions

We have presented a study of the relation between the interaction optical torque and plasmon hybridization in a dimer of twisted metal nanorods. Firstly, we have explained the physical description of plasmon hybridization in the dimer, which depends on not only the gap size but also the twisted angle between the nanorods. Then, we have demonstrated the interaction optical torque is highly associated with the plasmon hybridization by comparing the optical torques on the nanorods with the plasmon resonance spectra. This clearly shows the physical mechanism of the interaction optical torque based on the plasmon-hybridization-induced dipole momentum in the nanorods. The torque behaviors are decided by the relative phase between the coupled dipole moments of the two nanorods. In order to show how one can apply our findings to practical experiments and applications, we have obtained the rotational potential by using the calculated optical torques on Rod 2 in comparison with the Brownian motion energy. When the two-coupled dipoles are in phase and out of phase, i.e., hybridized anti-bonding and bonding modes, their rotations lead to mutually perpendicular and parallel alignments, respectively. In addition, we have discussed the experimental realization for the configuration of the simulation model. The optical alignments depending on the hybridized modes with the different resonance wavelength regions would dynamically control the plasmonic characteristics and functions, e.g., the EM field enhancement and chirality, through the nanoparticle configurations with the plasmon coupling. Thus, our findings will open a new route to all-optical active plasmonic and metamaterial devices.

Appendix

A.1 Calculation of optical torque by dipole approximation (DA) method

DA method is a widely utilized method to calculate the optical torque. A nanorod can be approximated as a dipole with polarizability of α. The induced dipole moment p of the nanorod illuminated by an arbitrary monochromatic incident electric field E can be expressed as

$${\textbf p}\textrm{ = }{\boldsymbol \alpha }{\textbf E ,}$$
Since the transverse polarizability of a nanorod is negligible [see Fig. 8(a)], we only consider the longitudinal polarizability αl. Figure 8(b) shows the real part and imaginary part of αl. Hence, with the incident field of y-polarized plane light E to propagate along –z-axis as shown in Fig. 7, the time-averaged optical torque on an isolated nanorod rotated with an angle of θ can be obtained by [46,47]
$${{\boldsymbol {N}}_{\textrm{isolated}}}\textrm{ = }\frac{\textrm{1}}{\textrm{2}}\textrm{Re[}{\textbf p}{ \times }{{\textbf E}^{\ast }}\textrm{] = } - \frac{\textrm{1}}{\textrm{4}}\textrm{Re[}{\alpha _l}\textrm{] sin2}\theta {|{\textbf E} |^\textrm{2}}{{\boldsymbol e}_\textrm{z}},$$
where ez is the unit vector along the z-axis.

 figure: Fig. 7.

Fig. 7. Schematic illustration of an isolated nanorod and twisted gold nanorods with gap size d and twisted angle θ illuminated by a y-polarized light. L = 120 nm, D = 40 nm.

Download Full Size | PDF

 figure: Fig. 8.

Fig. 8. (a) Plasmon resonance spectra of an isolated Au nanorod at longitudinal mode and transverse mode. (b) Longitudinal polarizability of the Au nanorod calculated by the finite element method. The length and diameter of the Au nanorods were 120 nm and 40 nm, respectively.

Download Full Size | PDF

Equation (7) shows a good agreement with the results calculated by the MST method (see Fig. 11). According to Eq. (7), the direction of optical torque is decided by the sign of the real part of αl. Under the situation of 0<θ<π/2, Re[αl] is positive in the longer wavelength region than the resonance peak wavelength. The phase difference between p and the E is in the range of 0 to π/2. The optical torque is negative. In the shorter wavelength region, Re[αl] is negative, that is, the phase difference between p and the incident electric field E is in the range of π/2 to π. The optical torque is positive. Additionally, when the phase difference between p and the incident electric field E is π/2, i.e., Re[αl] = 0, the optical torque is 0.

In order to calculate the optical torque acting on each nanorod in the dimer of twisted nanorods, we need to calculate the electric field on each nanorod by the dipole discrete approximation. Base on the dipole radiation, each component of the electric field on each nanorod can be expressed as

$${{\textbf E}_\textrm{1}}\textrm{ = }\left( \begin{array}{c} - \frac{{A\textrm{ + }{A^\textrm{2}}\textrm{exp[}ik\textrm{(}d\textrm{ + }D)\textrm{]}}}{{\textrm{1} - {A^\textrm{2}}\textrm{co}{\textrm{s}^\textrm{2}}\theta }}\textrm{sin}\theta \textrm{cos}\theta {\textrm{E}_\textrm{y}}\\ \textrm{ }\frac{{\textrm{exp[}ik\textrm{(}d\textrm{ + }D)\textrm{] + }A\textrm{co}{\textrm{s}^\textrm{2}}\theta }}{{\textrm{1} - {A^\textrm{2}}\textrm{co}{\textrm{s}^\textrm{2}}\theta }}{\textrm{E}_\textrm{y}}\textrm{ }\\ \textrm{ }\\ \textrm{ 0} \end{array} \right)\textrm{ ,}$$
$${{\textbf E}_\textrm{2}}\textrm{ = }\left( \begin{array}{c} \textrm{ 0}\\ \frac{{\textrm{1} + A\textrm{exp[}ik\textrm{(}d\textrm{ + }D)\textrm{]}}}{{\textrm{1} - {A^\textrm{2}}\textrm{co}{\textrm{s}^\textrm{2}}\theta }}{\textrm{E}_\textrm{y}}\\ \textrm{ 0} \end{array} \right),$$
$$A\textrm{ = }\frac{{{\alpha _l}\textrm{exp[}ik\textrm{(}d\textrm{ + }D)\textrm{]}}}{{\textrm{4}\pi \varepsilon \textrm{(}d + D\textrm{)}}}\textrm{ [}{k^\textrm{2}}\textrm{ + }\frac{{ik\textrm{(}d + D\textrm{)} - \textrm{1}}}{{{{\textrm{(}d + D\textrm{)}}^\textrm{2}}}}\textrm{],}$$
where θ is the twisted angle in the dimer, k is the wavenumber, ɛ is the permittivity of the surrounding medium, d is the gap size, and D is the diameter of nanorods. ∗ denotes the operation of conjugation. Ey is the y component of the incident light. Thus, the optical torques on Rod 1 and Rod 2 can be expressed as
$${{\boldsymbol N}_\textrm{1}}\textrm{ = }\frac{\textrm{1}}{\textrm{4}}\textrm{Re[}{\alpha _l}\frac{{{{\textrm{(}A\textrm{ + }{A^\textrm{2}}\textrm{exp[}ik\textrm{(}d + D\textrm{)])}}^{\ast }}\textrm{(exp[}ik\textrm{(}d + D\textrm{)] + }A\textrm{co}{\textrm{s}^\textrm{2}}\theta \textrm{)}}}{{{{|{\textrm{1} - {A^\textrm{2}}\textrm{co}{\textrm{s}^\textrm{2}}\theta } |}^\textrm{2}}}}\textrm{] sin2}\theta {|{\textbf E} |^\textrm{2}}\textrm{ }{{\boldsymbol e}_\textrm{z}},$$
$${{\boldsymbol N}_\textrm{2}}\textrm{ = } - \frac{\textrm{1}}{\textrm{4}}\textrm{Re[}{\alpha _l}\textrm{] (}\frac{{{{|{\textrm{1 + }A\textrm{exp[}ik\textrm{(}d\textrm{ + }D)\textrm{]}} |}^\textrm{2}}}}{{{{|{\textrm{1} - {A^\textrm{2}}\textrm{co}{\textrm{s}^\textrm{2}}\theta } |}^\textrm{2}}}}\textrm{)sin2}\theta {|{\textbf E} |^\textrm{2}}{{\boldsymbol e}_\textrm{z}}.$$
Though the DA method does not include the effect of the edges in the gap between the nanorods, it is a helpful method to qualitatively reveal the physics of the surface-plasmon-enhanced optical torque between the twisted nanorods. Figure 9 shows the optical torques on the Rod1 and Rod 2 calculated by Eqs. (11) and (12), respectively. The shapes of the curves show a similar tendency with the results calculated by the MST method in Fig. 4(d). For the optical torque on Rod 2, we can see the enhancement of optical torque is from the enhancement of the electric field illuminated on the nanorods by comparing the Eqs. (12) and (7). In the redshifted region, because the real part of polarizability changes slowly, as shown in Fig. 8(b), the enhancement of the electric field dominates the change of the optical torque. Thus, the peak of optical torque is corresponding to the plasmon coupling resonance. In the blueshifted region, the real part of polarizability varies drastically, determining the variation tendency of optical torque. At the peak wavelength of the plasmon resonance of an individual nanorod, we can find the optical torque is 0 in Fig. 9(b) because the real part of polarizability is 0. However, the DA method cannot completely describe the characteristics of the near field in the dimer, such as the polarization state distribution in the near field. Therefore, the optical torque shows different results from that in Fig. 4(d) (N2). For the optical torque on Rod 1, we can see the electric field on Rod 1 is not a linear polarized light from Eq. (8). In this case, the optical torque calculated by Eq. (11) is decided by not only the real part of polarizability but also the imaginary part of polarizability. Hence, in the blueshifted resonance region, the spectral behavior of optical torque is different from the optical torque on Rod 2. On the other hand, in the redshifted resonance region, it is similar to the optical torque on Rod 2 because the enhancement of the EM field dominates the variation of optical torque. The peak of optical torque is in accordance with the plasmon coupling resonance as well.

 figure: Fig. 9.

Fig. 9. Optical torques on Rod 1(a) and Rod 2 (b) calculated by the DA method as a function of wavelength in the dimer with 10 nm gap size.

Download Full Size | PDF

Figure 10 demonstrates the different results of optical torque between the MST method and DA method. When the gap size d is larger than ≈70 nm, the results from the two methods are in good agreement with each other. However, in the case of gap size smaller than ≈70 nm, the DA method is not accurate to calculate the optical torque in the dimer due to the emergence of plasmon coupling at the edges in the gap.

 figure: Fig. 10.

Fig. 10. Comparison of the optical torques acting on the whole structure (a), Rod 1(b) and Rod2 (c) calculated by MST and DA method as a function of gap size in the dimer with 710 nm wavelength and π/12 twisted angle.

Download Full Size | PDF

A.2 Optical torque on an isolated nanorod

 figure: Fig. 11.

Fig. 11. Optical torque on an isolated nanorod with different rotation angles θ. The solid line and triangle with different colors represent the torques calculated by the DA method and MST method, respectively.

Download Full Size | PDF

A.3 Experimental realization

Firstly, a nanorod array is embedded on the surface of a transparent substrate, e.g., silica, by etching trenches in the substrate. By controlling the depth of the trenches and the thickness of metal evaporated into the trenches, we can get a nanorod array a few nanometers below the surface of the substrate. The roughness of the surface can also be controlled within a few nanometers. Such fabrication has been realized by electron beam lithography and reactive ion etching in some paper already [48,49]. Then, we put nanorods dispersed in a solvent on the nanorod array. The dispersed nanorods would be aligned by controlling the incident light wavelength.

A.4 The optical force between twisted Au nanorods

As the optical force along the x-axis and y-axis relative to that along the z-axis can be neglected, we only consider the optical force along the z-axis. From Figs. 12 and 13, we can see the optical force shows a similar spectral behavior with the optical torque. The interaction optical force can be strongly enhanced by the plasmon coupling at the redshifted plasmon resonance, highly depending on the gap size and the twisted angle. At the redshifted resonance, a very strong attractive force can be generated between the two nanorods. At the blueshifted resonance, it is difficult to determine whether it is an attractive or repulsive force. However, due to the scattering force from the incident light, the total optical force on Rod 2 almost always keeps a negative value which is toward Rod 1. Thus, in an experiment that the bottom nanorod is fixed, the top nanorod can approach the bottom nanorod.

 figure: Fig. 12.

Fig. 12. Optical force acting on (a) Rod 1 and (b) Rod 2 with different gap sizes. The black line shows the optical force on an isolated Au nanorod. The twisted angle was fixed as π/12.

Download Full Size | PDF

 figure: Fig. 13.

Fig. 13. Optical force acting on (a) Rod1 and (b) Rod 2 with different twisted angles. The gap size was fixed as 1 nm.

Download Full Size | PDF

Funding

Precursory Research for Embryonic Science and Technology (JPMJPR15PA); Japan Society for the Promotion of Science (JP18K18992, JP19H02533, JP19H04670).

Acknowledgments

The authors are grateful to R. Fukuhara for many useful discussions. The authors acknowledge H. Sugimoto for primary calculations.

Disclosures

The authors declare no conflicts of interest.

References

1. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).

2. V. Giannini, A. I. Fernández-Domínguez, S. C. Heck, and S. A. Maier, “Plasmonic nanoantennas: fundamentals and their use in controlling the radiative properties of nanoemitters,” Chem. Rev. 111(6), 3888–3912 (2011). [CrossRef]  

3. T. W. Odom and G. C. Schatz, “Introduction to plasmonics,” Chem. Rev. 111(6), 3667–3668 (2011). [CrossRef]  

4. S. R. Pocock, X. Xiao, P. A. Huidobro, and V. Giannini, “Topological plasmonic chain with retardation and radiative effects,” ACS Photonics 5(6), 2271–2279 (2018). [CrossRef]  

5. L. Gunnarsson, T. Rindzevicius, J. Prikulis, B. Kasemo, M. Käll, S. Zou, and G. C. Schatz, “Confined plasmon in nanofabricated single silver particle pairs: Experimental observation of strong interparticle interaction,” J. Phys. Chem. B 109(3), 1079–1087 (2005). [CrossRef]  

6. W. Rechberger, A. Hohenau, A. Leitner, J. Krenn, B. Lamprecht, and F. Aussenegg, “Optical properties of two interacting gold nanoparticles,” Opt. Commun. 220(1-3), 137–141 (2003). [CrossRef]  

7. K.-H. Su, Q.-H. Wei, X. Zhang, J. J. Mock, D. R. Smith, and S. Schultz, “Interparticle coupling effects on plasmon resonances of nanogold particles,” Nano Lett. 3(8), 1087–1090 (2003). [CrossRef]  

8. E. Prodan, C. Radloff, N. J. Halas, and P. Nordlander, “A hybridization model for the plasmon response of complex nanostructures,” Science 302(5644), 419–422 (2003). [CrossRef]  

9. K. D. Osberg, N. Harris, T. Ozel, J. C. Ku, G. C. Schatz, and C. A. Mirkin, “Systematic study of antibonding modes in gold nanorod dimers and trimers,” Nano Lett. 14(12), 6949–6954 (2014). [CrossRef]  

10. P. Nordlander, C. Oubre, E. Prodan, K. Li, and M. Stockman, “Plasmon hybridization in nanoparticle dimers,” Nano Lett. 4(5), 899–903 (2004). [CrossRef]  

11. H. M. Abdulla, R. Thomas, and R. S. Swathi, “Overwhelming analogies between plasmon hybridization theory and molecular orbital theory revealed: The story of plasmonic heterodimers,” J. Phys. Chem. C 122(13), 7382–7388 (2018). [CrossRef]  

12. P. K. Jain, S. Eustis, and M. A. El-Sayed, “Plasmon coupling in nanorod assemblies: optical absorption, discrete dipole approximation simulation, and exciton-coupling model,” J. Phys. Chem. B 110(37), 18243–18253 (2006). [CrossRef]  

13. Z. Zhu, P. Yuan, S. Li, M. Garai, M. Hong, and Q.-H. Xu, “Plasmon-enhanced fluorescence in coupled nanostructures and applications in DNA detection,” ACS Appl. Bio Mater. 1(1), 118–124 (2018). [CrossRef]  

14. J. R. Lakowicz, K. Ray, M. Chowdhury, H. Szmacinski, Y. Fu, J. Zhang, and K. Nowaczyk, “Plasmon-controlled fluorescence: a new paradigm in fluorescence spectroscopy,” Analyst 133(10), 1308–1346 (2008). [CrossRef]  

15. X. Huang, I. H. El-Sayed, W. Qian, and M. A. El-Sayed, “Cancer cells assemble and align gold nanorods conjugated to antibodies to produce highly enhanced, sharp, and polarized surface Raman spectra: A potential cancer diagnostic marker,” Nano Lett. 7(6), 1591–1597 (2007). [CrossRef]  

16. B. Nikoobakht and M. A. El-Sayed, “Surface-enhanced Raman scattering studies on aggregated gold nanorods,” J. Phys. Chem. A 107(18), 3372–3378 (2003). [CrossRef]  

17. P. K. Jain and M. A. El-Sayed, “Noble metal nanoparticle pairs: effect of medium for enhanced nanosensing,” Nano Lett. 8(12), 4347–4352 (2008). [CrossRef]  

18. S. S. Aćimović, M. P. Kreuzer, M. U. González, and R. Quidant, “Plasmon near-field coupling in metal dimers as a step toward single-molecule sensing,” ACS Nano 3(5), 1231–1237 (2009). [CrossRef]  

19. J. Butet, P. F. Brevet, and O. J. F. Martin, “Optical second harmonic generation in plasmonic nanostructures: From fundamental principles to advanced applications,” ACS Nano 9(11), 10545–10562 (2015). [CrossRef]  

20. S. Kim, J. Jin, Y. J. Kim, I. Y. Park, Y. Kim, and S. W. Kim, “High-harmonic generation by resonant plasmon field enhancement,” Nature 453(7196), 757–760 (2008). [CrossRef]  

21. H. Xu and M. Käll, “Surface-plasmon-enhanced optical forces in silver nanoaggregates,” Phys. Rev. Lett. 89(24), 246802 (2002). [CrossRef]  

22. K. Wang, E. Schonbrun, and K. B. Crozier, “Propulsion of gold nanoparticles with surface plasmon polaritons: evidence of enhanced optical force from near-field coupling between gold particle and gold film,” Nano Lett. 9(7), 2623–2629 (2009). [CrossRef]  

23. Y. Tanaka, H. Yoshikawa, T. Itoh, and M. Ishikawa, “Surface enhanced Raman scattering from pseudoisocyanine on Ag nanoaggregates produced by optical trapping with a linearly polarized laser beam,” J. Phys. Chem. C 113(27), 11856–11860 (2009). [CrossRef]  

24. Y. Tanaka, H. Yoshikawa, T. Itoh, and M. Ishikawa, “Laser-induced self-assembly of silver nanoparticles via plasmonic interactions,” Opt. Express 17(21), 18760–18767 (2009). [CrossRef]  

25. A. J. Hallock, P. L. Redmond, and L. E. Brus, “Optical forces between metallic particles,” Proc. Natl. Acad. Sci. U. S. A. 102(5), 1280–1284 (2005). [CrossRef]  

26. A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu, “Observation of s single-beam gradient force optical trap for dielectric particles,” Opt. Lett. 11(5), 288–290 (1986). [CrossRef]  

27. D. G. Grier, “A revolution in optical manipulation,” Nature 424(6950), 810–816 (2003). [CrossRef]  

28. M. Hoshina, N. Yokoshi, H. Okamoto, and H. Ishihara, “Super-Resolution Trapping: A Nanoparticle Manipulation Using Nonlinear Optical Response,” ACS Photonics 5(2), 318–323 (2018). [CrossRef]  

29. H. J. Chen, Q. Ye, Y. W. Zhang, L. Shi, S. Y. Liu, J. Zi, and Z. F. Lin, “Reconfigurable lateral optical force achieved by selectively exciting plasmonic dark modes near Fano resonance,” Phys. Rev. A 96(2), 023809 (2017). [CrossRef]  

30. H. J. Chen, S. Y. Liu, J. Zi, and Z. F. Lin, “Fano resonance-induced negative optical scattering force on plasmonic nanoparticles,” ACS Nano 9(2), 1926–1935 (2015). [CrossRef]  

31. J. Chen, J. Ng, Z. F. Lin, and C. T. Chan, “Optical pulling force,” Nat. Photonics 5(9), 531–534 (2011). [CrossRef]  

32. S. B. Wang and C. T. Chan, “Lateral optical force on chiral particles near a surface,” Nat. Commun. 5(1), 3307 (2014). [CrossRef]  

33. Y. E. Lee, K. H. Fung, D. Jin, and N. X. Fang, “Optical torque from enhanced scattering by multipolar plasmonic resonance,” Nanophotonics 3(6), 343–440 (2014). [CrossRef]  

34. D. Gao, W. Ding, M. Nieto-Vesperinas, X. Ding, M. Rahman, T. Zhang, C. Lim, and C.-W. Qiu, “Optical manipulation from the microscale to the nanoscale: fundamentals, advances and prospects,” Light: Sci. Appl. 6(9), e17039 (2017). [CrossRef]  

35. M. Padgett and R. Bowman, “Tweezers with a twist,” Nat. Photonics 5(6), 343–348 (2011). [CrossRef]  

36. K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Extraordinary momentum and spin in evanescent waves,” Nat. Commun. 5(1), 3300 (2014). [CrossRef]  

37. A. Lehmuskero, R. Ogier, T. Gschneidtner, P. Johansson, and M. Käll, “Ultrafast spinning of gold nanoparticles in water using circularly polarized light,” Nano Lett. 13(7), 3129–3134 (2013). [CrossRef]  

38. R. A. Beth, “Mechanical detection and measurement of the angular momentum of light,” Phys. Rev. 50(2), 115–125 (1936). [CrossRef]  

39. A. Ashkin, “History of optical trapping and manipulation of small-neutral particle, atoms, and molecules,” IEEE J. Sel. Top. Quantum Electron. 6(6), 841–856 (2000). [CrossRef]  

40. M. Liu, T. Zentgraf, Y. Liu, G. Bartal, and X. Zhang, “Light-driven nanoscale plasmonic motors,” Nat. Nanotechnol. 5(8), 570–573 (2010). [CrossRef]  

41. M. E. J. Friese, T. A. Nieminen, N. R. Heckenberg, and H. Rubinsztein-Dunlop, “Optical alignment and spinning of laser-trapped microscopic particles,” Nature 394(6691), 348–350 (1998). [CrossRef]  

42. R. W. Bowman and M. J. Padgett, “Optical trapping and binding,” Rep. Prog. Phys. 76(2), 026401 (2013). [CrossRef]  

43. W. D. Phillips, “Nobel Lecture: Laser cooling and trapping of neutral atoms,” Rev. Mod. Phys. 70(3), 721–741 (1998). [CrossRef]  

44. S. Franke-Arnold, L. Allen, and M. Padgett, “Advances in optical angular momentum,” Laser Photonics Rev. 2(4), 299–313 (2008). [CrossRef]  

45. P. H. Jones, O. M. Maragò, and G. Volpe, Optical tweezers: Principles and applications (Cambridge University, 2015).

46. H. J. Chen, W. L. Lu, X. N. Yu, C. H. Xue, S. Y. Liu, and Z. F. Lin, “Optical torque on small chiral particles in generic optical fields,” Opt. Express 25(26), 32867–32878 (2017). [CrossRef]  

47. J.-W. Liaw, W.-J. Lo, and M.-K. Kuo, “Wavelength-dependent longitudinal polarizability of gold nanorod on optical torques,” Opt. Express 22(9), 10858–10867 (2014). [CrossRef]  

48. Y. Zhao, M. A. Belkin, and A. Alù, “Twisted optical metamaterials for planarized ultrathin broadband circular polarizers,” Nat. Commun. 3(1), 870 (2012). [CrossRef]  

49. Y. Zhao, A. N. Askarpour, L. Sun, J. Shi, X. Li, and A. Alù, “Chirality detection of enantiomers using twisted optical metamaterials,” Nat. Commun. 8(1), 14180 (2017). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (13)

Fig. 1.
Fig. 1. Schematic illustration of twisted gold nanorods with gap size d and twisted angle θ. The diameter (D) and length (L) of both nanorods are 40 nm and 120 nm, respectively.
Fig. 2.
Fig. 2. Plasmon resonance spectra of an isolated nanorod and its dimer: (a) gap size dependence (θ = π/12) and (b) twisted angle dependence (d = 1 nm). The solid lines, dashed lines, and dotted lines show the extinction, absorption and scattering cross sections, respectively. The vertical dashed lines show the resonance peak wavelength of the isolated nanorod.
Fig. 3.
Fig. 3. Plasmon hybridization diagram in nanorod dimer with twisted angle θ. The polarization direction of the incident light is along the longitudinal axis of the bottom nanorod.
Fig. 4.
Fig. 4. Optical torque acting on (a, c) the whole dimer and (b, d) on each nanorod: (a, b) gap size dependence (θ = π/12) and (c, d) twisted angle dependence (d = 1 nm). The dashed and dotdashed lines show the peak wavelengths at the redshifted resonance. The left- and right- side insets in Fig. 4(d) show the rotation directions of Rod 2 at the blueshifted and redshifted resonances, respectively.
Fig. 5.
Fig. 5. (a) Twisted angle dependence of electric field intensity enhancement at the gap center (d = 1 nm). The dot-dashed lines show the peak wavelengths at the redshifted resonance. (b, c) Distributions of the intensity enhancement at the resonance wavelength of 610 nm and 885 nm, respectively, with the twisted angle of π/12 at the gap center plane.
Fig. 6.
Fig. 6. Rotational potentials in units of kBT (T = 300K) for Rod 2 in the dimer (d = 1 nm) as a function of the twisted angle at different incident light wavelengths. Supercontinuum white lights with a uniform power over the spectrum from 500 nm to 645 nm (blueshifted resonance) and 645 nm to 1000 nm (redshifted resonance) are assumed to be the incident lights.
Fig. 7.
Fig. 7. Schematic illustration of an isolated nanorod and twisted gold nanorods with gap size d and twisted angle θ illuminated by a y-polarized light. L = 120 nm, D = 40 nm.
Fig. 8.
Fig. 8. (a) Plasmon resonance spectra of an isolated Au nanorod at longitudinal mode and transverse mode. (b) Longitudinal polarizability of the Au nanorod calculated by the finite element method. The length and diameter of the Au nanorods were 120 nm and 40 nm, respectively.
Fig. 9.
Fig. 9. Optical torques on Rod 1(a) and Rod 2 (b) calculated by the DA method as a function of wavelength in the dimer with 10 nm gap size.
Fig. 10.
Fig. 10. Comparison of the optical torques acting on the whole structure (a), Rod 1(b) and Rod2 (c) calculated by MST and DA method as a function of gap size in the dimer with 710 nm wavelength and π/12 twisted angle.
Fig. 11.
Fig. 11. Optical torque on an isolated nanorod with different rotation angles θ. The solid line and triangle with different colors represent the torques calculated by the DA method and MST method, respectively.
Fig. 12.
Fig. 12. Optical force acting on (a) Rod 1 and (b) Rod 2 with different gap sizes. The black line shows the optical force on an isolated Au nanorod. The twisted angle was fixed as π/12.
Fig. 13.
Fig. 13. Optical force acting on (a) Rod1 and (b) Rod 2 with different twisted angles. The gap size was fixed as 1 nm.

Equations (12)

Equations on this page are rendered with MathJax. Learn more.

d d t ( J mech  +  J field ) = S r ( T M × r ) n r d S r ,
T M = ε E E + 1 μ B B 1 2 ( ε E × E + 1 μ B × B ) I ,
N =   S r ( T ¯ M × r ) n r d S r ,
N i = 1 2 Re[ p × E ],
U p ( θ ) = θ 0 θ N d θ ,
p  =  α E ,
N isolated  =  1 2 Re[ p × E ] =  1 4 Re[ α l ] sin2 θ | E | 2 e z ,
E 1  =  ( A  +  A 2 exp[ i k ( d  +  D ) ] 1 A 2 co s 2 θ sin θ cos θ E y   exp[ i k ( d  +  D ) ] +  A co s 2 θ 1 A 2 co s 2 θ E y      0 )  ,
E 2  =  (  0 1 + A exp[ i k ( d  +  D ) ] 1 A 2 co s 2 θ E y  0 ) ,
A  =  α l exp[ i k ( d  +  D ) ] 4 π ε ( d + D )  [ k 2  +  i k ( d + D ) 1 ( d + D ) 2 ],
N 1  =  1 4 Re[ α l ( A  +  A 2 exp[ i k ( d + D )]) (exp[ i k ( d + D )] +  A co s 2 θ ) | 1 A 2 co s 2 θ | 2 ] sin2 θ | E | 2   e z ,
N 2  =  1 4 Re[ α l ] ( | 1 +  A exp[ i k ( d  +  D ) ] | 2 | 1 A 2 co s 2 θ | 2 )sin2 θ | E | 2 e z .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.