optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical...

7
Optically controlled stochastic jumps of individual gold nanorod rotary motors Downloaded from: https://research.chalmers.se, 2020-09-18 21:04 UTC Citation for the original published paper (version of record): Shao, L., Andrén, D., Jones, S. et al (2018) Optically controlled stochastic jumps of individual gold nanorod rotary motors PHYSICAL REVIEW B, 98(8) http://dx.doi.org/10.1103/PhysRevB.98.085404 N.B. When citing this work, cite the original published paper. research.chalmers.se offers the possibility of retrieving research publications produced at Chalmers University of Technology. It covers all kind of research output: articles, dissertations, conference papers, reports etc. since 2004. research.chalmers.se is administrated and maintained by Chalmers Library (article starts on next page)

Upload: others

Post on 22-Jul-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

Optically controlled stochastic jumps of individual gold nanorodrotary motors

Downloaded from: https://research.chalmers.se, 2020-09-18 21:04 UTC

Citation for the original published paper (version of record):Shao, L., Andrén, D., Jones, S. et al (2018)Optically controlled stochastic jumps of individual gold nanorod rotary motorsPHYSICAL REVIEW B, 98(8)http://dx.doi.org/10.1103/PhysRevB.98.085404

N.B. When citing this work, cite the original published paper.

research.chalmers.se offers the possibility of retrieving research publications produced at Chalmers University of Technology.It covers all kind of research output: articles, dissertations, conference papers, reports etc. since 2004.research.chalmers.se is administrated and maintained by Chalmers Library

(article starts on next page)

Page 2: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

PHYSICAL REVIEW B 98, 085404 (2018)

Optically controlled stochastic jumps of individual gold nanorod rotary motors

Lei Shao,1,* Daniel Andrén,1 Steven Jones,1 Peter Johansson,2,† and Mikael Käll1,‡1Department of Physics, Chalmers University of Technology, S-412 96 Göteborg, Sweden

2School of Science and Technology, Örebro University, S-701 82 Örebro, Sweden

(Received 1 May 2018; revised manuscript received 16 July 2018; published 2 August 2018)

Brownian microparticles diffusing in optical potential-energy landscapes constitute a generic test bed fornonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems,ranging from Josephson junctions to Carnot engines. Here we demonstrate that it is possible to scale down thisapproach to nanometric length scales by constructing a tilted washboard potential for the rotation of plasmonic goldnanorods. The potential depth and tilt can be precisely adjusted by modulating the light polarization. This allowsfor a gradual transition from continuous rotation to discrete stochastic jumps, which are found to follow Kramersdynamics in excellent agreement with stochastic simulations. The results widen the possibilities for fundamentalexperiments in statistical physics and provide insights into how to construct light-driven nanomachines andmultifunctional sensing elements.

DOI: 10.1103/PhysRevB.98.085404

I. INTRODUCTION

Laser tweezing is a powerful noninvasive tool to controland measure the movement of Brownian colloidal particlesfor applications in biology, physics, and chemistry [1–4]. Bytuning the light field distribution in the illuminated region, it ispossible to generate a wide variety of optical potential-energylandscapes. A thermally driven Brownian particle trappedin such a landscape represents an ideal model system forstudying a wide range of fundamental phenomena. Examplesto date include studies of chemical reactions [5,6], proteinfolding [7], thermodynamic relations [8], information flow [9],entropy production [10], and Kramers-type dynamics [11].By modulating optical potentials, it has even been possibleto realize optical Brownian ratchets [12] and single-particlemicroscopic Carnot engines [13]. Interested readers can refer to[14] to find more examples of using optically trapped colloidalparticles to build stochastic heat engines.

The tilted periodic “washboard” potential is a particu-larly important potential distribution realizable using opticaltweezers technology because it can be used as an archetypalnonequilibrium model of statistical physics in a variety ofsystems, such as the damped pendulum, ring-laser gyro-scope, Josephson junctions, superionic conduction, phase-locked loops, and charge-density-wave condensation [15–18].Despite the comprehensive theoretical discussion of Browniandynamics in a tilted washboard-type potential [16,19], it isstill of great interest to achieve convenient and precise shapecontrol of such optical potentials. Moreover, it is of genericinterest to be able to scale down the physical dimension of

*Also at Department of Physics, The Chinese University of HongKong, Shatin, Hong Kong SAR, China; [email protected]

†Also at Department of Physics, Chalmers University of Technol-ogy, S-412 96 Göteborg, Sweden.

[email protected]

the probe particle from the micron range to the nanoscalein order to approach length scales relevant to molecularinteractions and dynamics. In addition, nanoscale Brownianparticles have much shorter characteristic diffusion time thantheir micrometer counterparts and thus allow for faster andmore efficient sampling of thermodynamic transitions betweendifferent states.

In this paper, we employ elliptically polarized laser tweezersto create a tilted washboard rotational potential for trappingcolloidal gold nanorods [20]. The optical anisotropy andenhanced light-matter interaction of such particles, causedby plasmon resonances [21,22], result in extremely efficientoptical confinement and rotation performance [22,23]. Byadjusting the depth and tilt of the potential by controllingpolarization ellipticity, we successfully managed to switch therotational movement of a nanorod from ultrafast continuousspinning to discrete stochastic rotational jumps. Using bothexperiments and simulations, we further investigated the jumpdynamics at critical trapping polarizations and found that itquantitatively agrees with that predicted from Kramers theory[5,6]. The full control of Brownian rotation of plasmonicnanorods demonstrated here provides an additional freedomof nanomotor movement manipulation and holds great po-tential for future investigations of fundamental questions innonequilibrium thermodynamics. Our experimental configu-ration might also be useful for studies of molecular motors,optical Brownian ratchets, and optical torque wrenches forhigh-sensitivity biological experiments [24].

II. RESULTS

A. Construction of a rotational tilted washboard potential

An elliptically polarized plane wave with electric field E =E0[cos(ωt )x + cos(ωt + �φ )y] can be decomposed into onelinearly polarized and one circularly polarized component, EL

and EC, respectively (Fig. 1(a); a detailed analysis is providedin the Supplemental Material [25]). Once a gold nanorod is

2469-9950/2018/98(8)/085404(6) 085404-1 ©2018 American Physical Society

Page 3: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

SHAO, ANDRÉN, JONES, JOHANSSON, AND KÄLL PHYSICAL REVIEW B 98, 085404 (2018)

(a)

(d)

U

0 2 4 6Rotation angle ( )

U

(b)

(c)

Mc

kE

040

80

120

160

200

240 280

320

U+0

FIG. 1. Tilted washboard potential from elliptically polarizedlight. (a) An elliptically polarized plane wave can be decomposed intoone linear and one circular polarization component. (b) Schematic ofa gold nanorod in an optical trap. The nanorod is oriented with its longaxis perpendicular to the direction of incidence. (c) Nanorod potential-energy distribution induced by the linear polarization component of anelliptically polarized light field. The circular polarization componenttends to continuously rotate the rod with a torque of MC. (d) Thenanorod thus experiences a rotational tilted washboard potential-energy landscape as a function of orientation angle ϕ. The potentialbarrier height in the preferred rotation direction can be calculated to�U = 0.8kBT for parameters mimicking experimental data (nanorodwith dimensions of 65 × 147 nm2, λLaser = 830 nm, PLaser = 6 mW,T = 326 K) for the case when the elliptically polarized wave has�φ = 40◦. The barrier height in the opposite rotation direction is9.4kBT .

optically trapped [Fig. 1(b)], the linear component providesa restoring torque ML that tends to align it along the corre-sponding polarization direction while the circular componentinduces a torque MC that tends to spin the particle around thedirection of incidence to overcome the potential barrier formedby ML [Fig. 1(c)] [26]. Both torques are determined by angularmomentum transfer due to light absorption as well as scattering[22,27]. The response of the gold nanorod is approximately thatof a dipole with induced moment p = α · E. For an incidentwavelength close to the longitudinal plasmon resonance of theparticle, the polarizability tensor α is dominated by the longaxis component [20], which we denote α. The optical potentialexperienced by the nanorod then has a tilted washboard shape[Fig. 1(d)] according to

U (ϕ) = −∫ ϕ

0(−ML + MC)dϕ = −Aϕ + B sin2 ϕ, (1)

where A = 12 Re(α)E2

C, B = 12 Re(α)E2

L, and ϕ is the anglebetween the nanorod long axis and the linear polarizationcomponent EL [Fig. 1(d), inset]. It is easily shown that eachwell in U (ϕ) is surrounded by highly asymmetric barriers when�φ, the phase difference between the x and y field components,is below 45◦. For example, we find that the plasmonic nanorod

has to overcome a barrier height �U = 0.8kBT to rotatetowards the preferred direction, while the barrier in the oppositedirection is more than an order of magnitude higher forparameter settings mimicking our experimental conditions.The probability of rotational jumps in the “wrong” direction isthus very low. The barrier height �U can be varied by changingthe degree of polarization ellipticity via a change in �φ.

B. Transition from continuous rotation to stochastic jumpsof an individual gold nanorod

We studied the rotational dynamics of gold nanorods op-tically trapped in two dimensions against a cover glass in anoptical tweezers setup based on an 830-nm laser beam withtunable polarization ellipticity (see Methods in the Supple-mental Material [25]). The nanorods had an average size of(147 ± 10 nm) × (65 ± 5 nm) [Fig. 2(a)] and were preparedby a seed-mediated growth method [20,27]. The exemplarydark-field scattering spectrum of an individual trapped nanorod[Fig. 2(b)] shows one weak surface plasmon resonance ataround 550 nm and a strong mode at around 740 nm over-lapping the 830-nm trapping laser wavelength. The relativestrengths of the resonance peaks correspond to a polarizabilityalong the long axis that is more than an order of magnitudehigher than along the short axis at a wavelength of 830 nm, thusconfirming the assumption of an essentially one-dimensionalpolarizability tensor. The attractive and plasmon-enhancedoptical gradient force keeps the particle trapped and alignedin the laser focus xy plane, while the Coulomb repulsion fromthe cover glass and the radiation pressure prevent the particlefrom escaping the trap along the z axis [27].

We first tracked the rotational dynamics of a trappednanorod by analyzing the backscattered laser light I P

sca fromthe particle using polarization-selective detection in whicha polarizer oriented perpendicular to the linear polarizationcomponent EL of the trapping beam has been placed in frontof the detector. The nanorod angle variations are thus convertedto fluctuations in I P

sca because of the nanorods’ highly polarizedscattering properties [27]. I P

sca(t ) shows a periodic oscillationwith superimposed fluctuations due to rotational Brownian mo-tion for the case of an almost circularly polarized trapping field[�φ = 88◦; Fig. 2(c), black trace]. The corresponding auto-correlation function C(τ ) of I P

sca(t ) [Fig. 2(d), black trace] canbe analyzed using C(τ ) = I 2

0 + 0.5I 21 exp(−τ/τ0) cos(4πf τ )

[27], which yields the nanorod average rotation frequency asf = 2460 ± 20 Hz and the autocorrelation decay time as τ0 =103 ± 2 μs. The nanorod ceased to rotate when we switchedthe laser polarization to elliptical (�φ = 37◦), as is evidentfrom the lack of a well-defined periodicity in the measuredI P

sca(t ) and C(τ ) [Figs. 2(f) and 2(g), black traces]. However,the recorded intensity trace nevertheless exhibits distinct occa-sional burst. We interpret these features as being due to well-defined but stochastic thermal jumps in nanorod orientation.

Next, stochastic simulations were performed to gain furtherinsight into the rotation process. The Brownian dynamics of ananorod trapped in the tilted washboard potential U (ϕ) can besimulated using the equation of motion [16,27]:

J ϕ = −γrϕ − cos ϕ sin ϕRe(α)E2L + 1

2 Re(α)E2C + ξ (t ).

(2)

085404-2

Page 4: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

OPTICALLY CONTROLLED STOCHASTIC JUMPS OF … PHYSICAL REVIEW B 98, 085404 (2018)

(a)

Sca

tterin

g (a

.u.)

800700600500Wavelength (nm)

(b)

experimentsimulation(i)

02468

(k T )B

U

3

2

1

0Freq

uenc

y (k

Hz)

806040200 o( )

AC

F

10-5 10-4 10-3

Lag time (s)

4

2

0Inte

nsity

(a.u

.)

3210Time (ms)

90

60

30

020151050

Time (ms)

Rot

atio

n (

)

(e)(c) (d)

9

6

3

020151050

Time (ms)151050

Time (ms)

4

2

0Inte

nsity

(a.u

.)

)h()f(

10-5 10-4 10-3

Lag time (s)

AC

F

(g)

** *

* **

***

** ** * * * **

1.6 k TBU~Rot

atio

n (

)

0 k TBU~

FIG. 2. Transition from continuous rotation to discrete jumps of gold nanorods. (a) Scanning electron microscopy image of the nanorods(scale bar is 200 nm). (b) Scattering spectrum of a trapped nanorod. The strong peak at ∼740 nm is caused by the long-axis surfaceplasmon resonance of the nanorod. (c)–(h) Measured (black) and simulated (red) rotational dynamics of (c)–(e) the rod undergoing continuousrotation in the presence of an almost perfectly circularly polarized laser field (�φ = 88◦) and (f)–(h) discrete rotational jumps due to anelliptically polarized field (�φ = 37◦). (e) and (h) show the nanorod orientation angle ϕ versus time, while (c) and (d) and (f) and (g)show the corresponding scattering intensity time traces and intensity autocorrelation functions (ACFs), respectively. The data are based onmeasurements and simulations of cross-polarized backscattering from the nanorod. The blue stars in (f) and (h) mark individual jumps. (i)Polarization-dependent rotation of a nanorod. We showed both measured (black) and calculated (red) rotational frequency of a nanorod as thelaser polarization continuously varies from almost circular to linear. The blue curve indicates the barrier height at varying �φ. In the highlightedarea (30◦ < �φ < 45◦), where the potential barrier �U ∈ (0, 4kBT ), we observe the nanorod undergoing a transition from continuous rotationto discrete jumps. The nanorod stops continuous rotation when �φ decreases to around 40◦, where an effective rotational potential barrier�U ≈ kBT .

Here J is the nanorod moment of inertia, and the first threeterms on the right-hand side represent, respectively, a viscousdamping torque, characterized by a rotational friction coef-ficient γr; the restoring torque due to the linear polarizationcomponent EL; and the driving torque due to the circular po-larization component EC. The last term represents a stationaryGaussian noise torque with zero mean and autocorrelationfunction 〈ξ (t )ξ (0)〉 = 2γrkBTrδ(t ), where Tr is the effectivetemperature for rotational Brownian motion [28]. The temporalvariation in nanorod orientation ϕ(t ) obtained from Eq. (2) can,in turn, be used to calculate I P

sca(t ) and C(τ ) for comparisonwith experiments.

Figures 2(c)–2(h) (red traces) show simulation resultsfor almost circular (�φ = 88◦) and elliptical (�φ = 37◦)polarization using simulation parameters selected to matchexperimental conditions, including a fixed Tr = 320 K esti-mated [27,28] from the experimental τ0 (see the SupplementalMaterial for details [25]). For the circular polarization case, thesimulated ϕ(t ) evolves continuously [Fig. 2(e)], correspondingto continuous rotation, and the calculated C(τ ) yields f =2643 ± 7 Hz and τ0 = 100 ± 1 μs, in excellent agreementwith the experimental results. For the elliptical polarizationcase, ϕ(t ) instead exhibits a staircase behavior [Fig. 2(h)]corresponding to discrete and random π jumps in one directiongiven by the driving torque, separated by periods of almostfixed alignment along the linear polarization direction. The re-sulting intensity trace and autocorrelation function are again in

good agreement with the experimental observations [Figs. 2(f)and 2(g)]. Thus, the simulations confirm that thermal agitationoccasionally forces the nanorod to jump out of the local minimaof the washboard potential to the next, lower potential well,resulting in discrete intensity bursts in I P

sca that can be trackedexperimentally.

We further varied the polarization state of the trappinglaser continuously. Both experiments and simulations showedthat the gold nanorod rotation becomes increasingly slow asthe polarization becomes increasingly elliptical. When �φ

decreases to ∼40◦, the oscillating feature in the intensityautocorrelation function disappears (Fig. 2(i); more details canbe found in the Supplemental Material [25]), suggesting thata barrier exists and stops the nanorod’s continuous rotation.However, the nanorod undergoes stochastic rotational jumpsfrom time to time, and we can still calculate an effectiverotation frequency by counting the number of discrete jumps.When the polarization becomes even more elliptical, the barrierin the rotation potential becomes high enough to keep therod aligned with the major axis of the polarization ellipse.The nanorod thus exhibits a “rotation” frequency asymptotictowards zero. From simulation, we further observed that theeffective rotational diffusion of the nanorod varies as the laserpolarization changes, similar to the reported result of trans-lational diffusion of a Brownian particle on tilted washboardpotentials [29]. Interested readers can find more discussion inthe Supplemental Material [25].

085404-3

Page 5: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

SHAO, ANDRÉN, JONES, JOHANSSON, AND KÄLL PHYSICAL REVIEW B 98, 085404 (2018)

(d)

10

50

100

150

0 0.2 0.4 0.6 0.8Transit time (ms)

Num

ber o

f flip

s

0

100

200

300

0 5 10 15 20Flipping steps in 5 ms

Pro

babi

lity

0 5 10 15 200

10

20

30

40

Flipping steps in 5 ms

Pro

babi

lity

(c)

0 0.2 0.4 0.6 0.8 10

50

100

150

200

Transit time (ms)

Num

ber o

f flip

s

(a) (b)

200250

experiment simulation

experiment simulation

T

FIG. 3. Stochastic rotational jump dynamics of a gold nanorod.The gold nanorod was trapped using an elliptically polarized laserbeam with �φ = 37◦ and a power of 6 mW. (a) and (b) Experimen-tally measured and simulation-calculated probability distributions ofnanorod flips in a fixed time interval of 5 ms and (c) and (d) transittimes of such a jump. The nanorod rotational effective Browniantemperature Tr was set at 360 K in simulation to achieve goodagreement with experiments. The laser trap forms a tilted washboardrotation potential with a barrier �U = 1.4kBTr, and near a localmaximum the potential landscape can be modeled with an invertedharmonic potential [inset in (a)]. The rod was agitated thermally toovercome this barrier to jump by an angle of π , and each jump processwas recorded by one scattering intensity peak shown in Fig. 2(f). Thenumbers of jumps follow Poisson distribution [blue curves in (a) and(b)]. The full distributions of transit times are well fitted by the formuladerived from Kramers theory [blue curves in (c) and (d)], with thecoefficients of determination R2 = 0.979 and 0.994, respectively.

C. Stochastic jump dynamics of gold nanorods trappedby elliptical polarization

To test quantitatively the physical attributes of nanorodstochastic jumps, including their rate and transit time, wefurther examined the measured scattering signals and thesimulation results to discern further details. Statistics for boththe experiment and simulation results revealed that the numberof nanorod flips in a certain time interval X follows a Poissonprobability distribution P (X = S) = e−λλS/S! [Figs. 3(a) and3(b)], with the Poisson mean λ determined by the potentialbarrier height relative to kBTr. When the time interval is setat 5 ms, the fitting-obtained λ are 4.3 ± 0.3 for experimentand 4.7 ± 0.1 for simulation (laser power 6 mW, �φ = 37◦),respectively.

Additional information can be obtained through studyingthe duration of each individual stochastic jump, as the averagevalue and the variability in transit times reflect kinetics and thefundamentally statistical nature of the stochastic jump process.The transit time τT is much shorter than the first-passage timedefined as how long it takes for the nanorod to rotate byπ . This is very different from the case when the nanorodundergoes continuous rotation (more details can be foundin the Supplemental Material [25]). τT was found to vary

widely, from less than 80 to over 600 μs [Figs. 3(c) and3(d)], with average values 〈τT〉 = 158 ± 114 μs (experiment)and 117 ± 47 μs (simulation). Additionally, both experimentaland simulation results revealed that the broadly distributedτT has a peak at around 120 μs and a long exponential tail[Figs. 3(c) and 3(d)]. This behavior is similar to that expectedfor transit across harmonic barriers in the high-barrier limit(�U � kBT ) in the Kramers regime [5,6,30]. Specifically, inour case, the potential landscape near a local maximum canbe modeled with an inverted harmonic potential with a “springconstant” κb, V (ϕ) ≈ Vmax − κb(ϕ − ϕmax)2/2 [schematic inFig. 3(a), inset]. When the transition region is from (ϕmax − ϕ0)to (ϕmax + ϕ0), the barrier height �U = κbϕ

20/2, and the transit

time τT has a distribution P (τT). For the one-dimensionaldiffusion model determined by J ϕ = −γrϕ − V ′(ϕ) + ξ (t ),P (τT) is predicted to have the form [30,31]

P (τT) = ωK√

�U/(kBT )

1 − erf[√

�U/(kBT )]

× exp[−�U coth(ωKτT/2)/(kBT )]

sinh(ωKτT/2)√

2π sinh(ωKτT). (3)

The distribution in Eq. (3) decays exponentially for large τT

as P (τT) ≈ 2ωK[�U/(kBT )]exp(−ωKτT). The parameter ωK

sets the timescale for decay away from states near the top ofthe barrier, ωK = κb/γr.

In both experiment and simulation, P (τT) are well fittedby Eq. (3) [Figs. 3(c) and 3(d)]. The barrier heights �U

returned by the fit are (1.10 ± 0.30)kBT (experiment) and(1.36 ± 0.08)kBT (simulation), both in good agreement withthe value calculated from the laser polarization according toEq. (1): �U = 1.4kBTr. The values of fitting-obtained ωK

are 1.13 ± 0.17 × 104 s−1 (experiment) and 1.02 ± 0.03 ×104 s−1 (simulation). Given that the rotational diffusion con-stant Dr is determined by Dr = kBT/γr, we can write ωK aswell as the average transit time in terms of this quantity through

ωK = κb/γr = Drκb/(kBT ), (4a)

〈τT〉 = ln[2eγ �U/(kBT )]/ωK, (4b)

where γ is Euler’s constant [30]. Given that κb can bedetermined by fitting the energy landscape to be κb = (3.4 ±0.1)kBT/rad2 at �φ = 37◦, we can calculate the rotationaldiffusion coefficient Dr from ωK and 〈τT〉 according to Eqs. (4).For the simulation result, Dr calculated from ωK is 3.0 ±0.1 × 103 s−1, close to Dr = 4.0 ± 1.6 × 103 s−1 calculatedfrom 〈τT〉. In experiment, Dr is calculated from ωK and 〈τT〉to be 3.3 ± 0.5 × 103 and 3.0 ± 2.2 × 103 s−1, respectively,which are also in good agreement with each other. The valuesof Dr calculated from measured ωK and 〈τT〉 are close tothe result directly calculated by modeling the nanorod as aprolate ellipsoid in water (Dr = 4.6 × 103 s−1), validating theKramers description of the nanorod rotational jump transition.

Furthermore, the rates and transit times of the nanorodstochastic jumps are highly dependent on the temperatureand viscosity of the local nanoenvironment. If we artificiallyincrease Tr in simulation, the calculated Poisson distribu-tion means λ and the average transit time 〈τT〉 exhibit arapid increase and decrease [Figs. 4(a) and 4(b)], respec-tively. λ, indicating the transition rate, follows an exponential

085404-4

Page 6: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

OPTICALLY CONTROLLED STOCHASTIC JUMPS OF … PHYSICAL REVIEW B 98, 085404 (2018)

Tran

sit t

ime

(100

s)

450400350300Temperature (K)

)b()a(10

5

0450400350300

Temperature (K)

=A+Be-C U/k TB

4

210

3 =a+be-c(T-T*)

3

2

1

1.81.51.2

5

4

3

21.81.51.2 Tr

ansi

t tim

e (1

00

s)

/ (water) / (water)

(c) (d)

FIG. 4. Temperature- and medium-viscosity-dependent nanorodstochastic jump dynamics. (a) Rotational effective Brownian tem-perature Tr-dependent average number of flips in a time interval of5 ms and (b) transit time calculated from simulation (red dots). Theblue curve in (a) and the red curve in (b) are fitting results, withthe coefficients of determination R2 = 0.975 and 0.998. (c) Mediumviscosity η-dependent average number of flips and (d) transit timecalculated from simulation (red dots). The blue curve in (c) and thered curve in (d) are exponential and linear fitting results, with thecoefficients of determination R2 = 0.999 and 0.968.

(Boltzmann) dependence on the energy barrier height �U

[Fig. 4(a)]. 〈τT〉 decreases exponentially with Tr [Fig. 4(b)].λ and 〈τT〉 are also very sensitive to the medium viscosity η.As η increases, λ decreases exponentially, and 〈τT〉 increaseslinearly [Figs. 4(c) and 4(d)], according to our simulationresults. The sensitive temperature and viscosity dependenceof the nanorod stochastic jump dynamics implies that the goldnanorod manipulated by elliptical polarization can work as asensing element to probe the local temperature and viscosityin solution.

III. DISCUSSION

We have shown the construction of tilted washboard ro-tational potentials for optically trapped Brownian plasmonicgold nanorod motors by rather simple means, utilizing ellip-tical polarizations. The gold nanorods were modulated fromcontinuous rotation to discrete jumps by simply adjustingthe polarization state of the trapping laser. In addition, we

studied individual stochastic jump processes of the nanorodat critical laser polarization, finding that the jump dynamics isin good agreement with that predicted by Kramers theory. Ourmeasurement results can be directly used to help understandmechanisms in molecular motors [32] and rotating dipoles inexternal fields [33].

The plasmonic nanorod trapped by elliptical polarizationis a simple optical analogy to many other physical systems.It provides a powerful tool for investigating fundamentalquestions where the problem of Brownian motion in tiltedperiodic potentials arises [15–18]. The colloidal nanoparticletrapped in solution works in an overdamped regime; one canextend this to an underdamped regime by trapping plasmonicnanoparticles in air or in vacuum [11,34]. The plasmonicnanorod optically trapped with elliptical polarization thusgenerates a universal model system for the nonequilibriumthermodynamics problem of Brownian diffusion over periodicbarriers. The small size (∼100 nm) and short characteristictimescale (∼100 μs) of the Brownian nanorod allow for faststatistical investigations. Experimental parameters can be con-trolled and varied in situ easily in an optical way, significantlyreducing the complexity, incompatibility, and inadaptability ofother physical systems conventionally employed. As we haveshown, a very simple and well-known theoretical model canbe utilized to capture the system dynamics [16,19]. This, inturn, means that it is also straightforward and robust to extractparameters by model fitting. As a result, we believe that ourstudy facilitates the use of Brownian plasmonic nanoparticlesas new probes to study fundamental issues with broad interest,such as giant acceleration of particle diffusion [29], connectionbetween statistical physics and information theory [9,35], andhydrodynamic synchronization [36].

In addition, we have realized the full rotational control of thelight-driven gold nanorod motors. The elliptical-polarization-trapped gold nanorod can also work as a ratchet that harvestsoverdamped Brownian noise and rectifies the Brownian motionat thermal nonequilibrium [12]. Moreover, by combining thestructure Brownian dynamics analysis, which can probe thelocal viscosity and temperature, and the plasmonic molecularanalysis techniques such as refractometric sensing [37] andsurface-enhanced Raman scattering [38], the optical-potential-controlled gold nanorod further becomes a multifunctionalsensing platform to probe different characteristics of the localnanoenvironment [39,40].

ACKNOWLEDGMENTS

The authors thank N. Odebo Länk for help with thefinite-difference time-domain simulations. Support and advicefrom Prof. G. Volpe and Prof. A. Isacsson is gratefullyacknowledged. This work was supported by the Knut and AliceWallenberg Foundation.

[1] A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu,Observation of a single-beam gradient force optical trap fordielectric particles, Opt. Lett. 11, 288 (1986).

[2] T. Yamane A. Ashkin, and J. M. Dziedzic, Optical trapping andmanipulation of single cells using infrared laser beams, Nature(London) 330, 769 (1987).

085404-5

Page 7: Optically controlled stochastic jumps of individual gold ... · nonequilibrium statistical thermodynamics and have been used to emulate a wide variety of physical systems, ... dynamics

SHAO, ANDRÉN, JONES, JOHANSSON, AND KÄLL PHYSICAL REVIEW B 98, 085404 (2018)

[3] A Ashkin and J. M. Dziedzic, Optical trapping and manipulationof viruses and bacteria, Science 235, 1517 (1987).

[4] J. R. Moffitt, Y. R. Chemla, S. B. Smith, and C. Bustamante,Recent advances in optical tweezers, Annu. Rev. Biochem. 77,205 (2008).

[5] H. A. Kramers, Brownian motion in a field of force and thediffusion model of chemical reactions, Physica (Amsterdam) 7,284 (1940).

[6] P. Hänggi, P. Talkner, and M. Borkovec, Reaction-rate theory:Fifty years after Kramers, Rev. Mod. Phys. 62, 251 (1990).

[7] K. Neupane, D. A. N. Foster, D. R Dee, H. Yu, F. Wang, andM. T. Woodside, Direct observation of transition paths during thefolding of proteins and nucleic acids, Science 352, 239 (2016).

[8] G. M. Wang, E. M Sevick, E. Mittag, D. J. Searles, and D. J.Evans, Experimental Demonstration of Violations of the SecondLaw of Thermodynamics for Small Systems and Short TimeScales, Phys. Rev. Lett. 89, 050601 (2002).

[9] A. Bérut, A. Arakelyan, A. Petrosyan, S. Ciliberto, R. Dillen-schneider, and E. Lutz, Experimental verification of Landauer’sprinciple linking information and thermodynamics, Nature(London) 483, 187 (2012).

[10] C. Tietz, S. Schuler, T. Speck, U. Seifert, and J. Wrachtrup,Measurement of Stochastic Entropy Production, Phys. Rev. Lett.97, 050602 (2006).

[11] L. Rondin, J. Gieseler, F. Ricci, R. Quidant, C. Dellago, andL. Novotny, Direct measurement of Kramers turnover with alevitated nanoparticle, Nat. Nanotechnol. 12, 1130 (2017).

[12] S.-H. Wu, N. Huang, E. Jaquay, and M. L. Povinelli, Near-field,on-chip optical Brownian ratchets, Nano Lett. 16, 5261 (2016).

[13] I. A. Martínez, É. Roldán, L. Dinis, D. Petrov, J. M. R. Parrondo,and R. A. Rica, Brownian Carnot engine, Nat. Phys. 12, 67(2016).

[14] I. A. Martinez, E. Roldan, L. Dinis, and R. A. Rica, Colloidalheat engines: A review, Soft Matter 13, 22 (2017).

[15] Advances in Chemical Physics, edited by S. A. Rice and A. R.Dinner (Wiley, Hoboken, NJ, 2013), Vol. 153.

[16] H. Risken, Fokker-Planck equation, in The Fokker-Planck Equa-tion: Methods of Solution and Applications (Springer, Berlin,1996), pp. 63–95.

[17] P. Fulde, L. Pietronero, W. R. Schneider, and S. Strässler,Problem of Brownian Motion in a Periodic Potential, Phys. Rev.Lett. 35, 1776 (1975).

[18] R. L. Stratonovich, Topics in the Theory of Random Noise (CRCPress, Boca Raton, 1967), Vol. 2.

[19] W. T. Coffey and Y. P. Kalmykov, The Langevin Equation: WithApplications to Stochastic Problems in Physics, Chemistry andElectrical Engineering (World Scientific, Singapore, 2004).

[20] H. Chen, L. Shao, Q. Li, and J. Wang, Gold nanorods and theirplasmonic properties, Chem. Soc. Rev. 42, 2679 (2013).

[21] A. Lehmuskero, P. Johansson, H. Rubinsztein-Dunlop, L. Tong,and M. Käll, Laser trapping of colloidal metal nanoparticles,ACS Nano 9, 3453 (2015).

[22] L. Shao and M. Käll, Light-driven rotation of plasmonicnanomotors, Adv. Funct. Mater. 28, 1706272 (2018).

[23] M. E. J. Friese, T. A. Nieminen, N. R. Heckenberg, andH. Rubinsztein-Dunlop, Optical torque controlled by ellipticalpolarization, Opt. Lett. 23, 1 (1998).

[24] F. Pedaci, Z. Huang, M. Van Oene, S. Barland, and N. H. Dekker,Excitable particles in an optical torque wrench, Nat. Phys. 7, 259(2011).

[25] See Supplemental Material at http://link.aps.org/supplemental/10.1103/PhysRevB.98.085404, which includes Refs. [41–43],for more details and discussion related to this paper.

[26] P. V. Ruijgrok, N. R. Verhart, P. Zijlstra, A. L. Tchebotareva,and M. Orrit, Brownian Fluctuations and Heating of an Op-tically Aligned Gold Nanorod, Phys. Rev. Lett. 107, 037401(2011).

[27] L. Shao, Z.-J. Yang, D. Andrén, P. Johansson, and M. Käll, Goldnanorod rotary motors driven by resonant light scattering, ACSNano 9, 12542 (2015).

[28] F. Hajizadeh, L. Shao, D. Andrén, P. Johansson, H. Rubinsztein-Dunlop, and M. Käll, Brownian fluctuations of an opticallyrotated nanorod, Optica 4, 746 (2017).

[29] P. Reimann, C. Van den Broeck, H. Linke, P. Hänggi, J. M. Rubi,and A. Pérez-Madrid, Giant Acceleration of Free Diffusion ByUse of Tilted Periodic Potentials, Phys. Rev. Lett. 87, 010602(2001).

[30] S. Chaudhury and D. E. Makarov, A harmonic transition stateapproximation for the duration of reactive events in com-plex molecular rearrangements, J. Chem. Phys. 133, 034118(2010).

[31] B. W. Zhang, D. Jasnow, and D. M. Zuckerman, Transition-eventdurations in one-dimensional activated processes, J. Chem. Phys.126, 074504 (2007).

[32] J. Michl and E. C. H. Sykes, Molecular rotors and motors: Recentadvances and future challenges, ACS Nano 3, 1042 (2009).

[33] E. Praestgaard and N. G. Van Kampen, A model for rotationalrelaxation and resonance, Molec. Phys. 43, 33 (1981).

[34] L. Jauffred, S. M.-R. Taheri, R. Schmitt, H. Linke, and L. B.Oddershede, Optical trapping of gold nanoparticles in air, NanoLett. 15, 4713 (2015).

[35] J. M. R. Parrondo, J. M. Horowitz, and T. Sagawa, Thermody-namics of information, Nat. Phys. 11, 131 (2015).

[36] N. Koumakis and R. Di Leonardo, Stochastic HydrodynamicSynchronization in Rotating Energy Landscapes, Phys. Rev.Lett. 110, 174103 (2013).

[37] P. Zijlstra, P. M. R. Paulo, and M. Orrit, Optical detectionof single non-absorbing molecules using the surface plasmonresonance of a gold nanorod, Nat. Nanotechnol. 7, 379 (2012).

[38] H. Xu, E. J. Bjerneld, M. Käll, and L. Börjesson, Spectroscopyof Single Hemoglobin Molecules by Surface Enhanced RamanScattering, Phys. Rev. Lett. 83, 4357 (1999).

[39] D. Andrén, L. Shao, N. Odebo Länk, S. S. Acimovic, P.Johansson, and M. Käll, Probing photothermal effects on op-tically trapped gold nanorods by simultaneous plasmon spec-troscopy and Brownian dynamics analysis, ACS Nano 11, 10053(2017).

[40] H. Šípová, L. Shao, N. Odebo Länk, D. Andrén, and M.Käll, Photothermal DNA release from laser-tweezed individualgold nanomotors driven by photon angular momentum, ACSPhotonics 5, 2168 (2018).

[41] G. Falasco, M. V. Gnann, D. Rings, and K. Kroy, Effectivetemperatures of hot Brownian motion, Phys. Rev. E 90, 032131(2014).

[42] S.-H. Lee and D. G. Grier, Giant Colloidal Diffusivity onCorrugated Optical Vortices, Phys. Rev. Lett. 96, 190601 (2006).

[43] R. Hayashi, K. Sasaki, S. Nakamura, S. Kudo, Y. Inoue, H. Noji,and K. Hayashi, Giant Acceleration of Diffusion Observed ina Single-Molecule Experiment on F1-ATPase, Phys. Rev. Lett.114, 248101 (2015).

085404-6