Demonstration of Ultra Low Dissipation Optomechanical Resonators on a Chip

Total Page:16

File Type:pdf, Size:1020Kb

Demonstration of Ultra Low Dissipation Optomechanical Resonators on a Chip Demonstration of Ultra Low Dissipation Optomechanical Resonators on a Chip G. Anetsberger, R. Rivi`ere,A. Schliesser, O. Arcizet & T. J. Kippenberg∗ Max-Planck-Institut f¨urQuantenoptik, Hans-Kopfermann-Str.1, 85748 Garching Cavity-enhanced radiation-pressure coupling of optical and mechanical degrees of freedom gives rise to a range of optomechanical phenomena, in particular providing a route to the quantum regime of mesoscopic mechanical oscillators. A prime challenge in cavity optomechanics has however been to realize systems which simultaneously maximize optical finesse and mechanical quality. Here we demonstrate for the first time independent control over both mechanical and optical degree of freedom within one and the same on-chip resonator. The first direct observation of mechanical normal mode coupling in a micromechanical system allows for a quantitative understanding of mechanical dissipation. Subsequent optimization of the resonator geometry enables intrinsic material loss limited mechanical Q-factors, rivalling the best values reported in the high MHz frequency range, while simultaneously preserving the resonators' ultra-high optical finesse. Besides manifesting a complete understanding of mechanical dissipation in microresonator based optomechanical systems, our results provide an ideal setting for cavity optomechanics. Over the past years it has become experimentally pos- of both is important for applications such as low loss, sible to study the coupling of optical and mechanical narrow-band \photonic clocks"[2, 27] and indispensable modes via cavity enhanced radiation pressure, which for fundamental studies aiming at approaching and de- gives rise to a diverse set of long anticipated optome- tecting quantized motion in mesoscopic optomechanical chanical phenomena[1] such as radiation pressure driven systems[11, 12, 28]. oscillations[2] and, as demonstrated in 2006, dynamic Here we show for the first time independent control over backaction cooling[3, 4, 5, 6, 7]. Moreover, this coupling both optical and mechanical degree of freedom in one can be exploited to perform highly sensitive measure- and the same microscale optomechanical resonator. The ments of displacement[8, 9] which may enable the obser- demonstration of mechanical normal mode coupling[29] vation of radiation pressure quantum backaction[10] or within a micromechanical device and the concomitant ge- related phenomena. Major goals in the emerging field ometry dependence of clamping losses allows a quantita- of cavity-optomechanics[11, 12], such as ground-state tive understanding of mechanical dissipation. We demon- cooling[13, 14] necessitate high optical finesse and high strate monolithic spoke-supported silica resonators com- mechanical quality factors at mechanical oscillation fre- prising a toroidal boundary which allows ultra-high op- quencies exceeding the optical cavity's linewidth. While tical finesse (> 106)[24] rivalling the best values ob- recently impressive progress has been made in creating tained in Fabry-Perot cavities[25]. Independent con- experimental settings in which radiation pressure effects trol of their mechanical properties leads to strongly re- can be studied[3, 4, 5, 15, 16, 17], a prime challenge still duced clamping losses allowing for unprecedented me- concerns attaining simultaneously high optical finesse chanical quality factors (e.g. 80,000 at 38 MHz) { ri- and high mechanical Q-factors. Nearly all approaches so valling the best published values of strained silicon nitride far have combined traditional cm-sized optical elements nanoresonators[19, 20] as well as radial contour-mode (mirrors) with micro- or nanoscale mechanical oscillators disk resonators[21] at similarly high frequencies which which simultaneously act as mirrors[3, 4, 15, 16] (or as allow entering the resolved sideband regime[30, 31] using dispersive element[17]). Other approaches have used the state-of-the-art optical cavities. It is shown that the mea- intrinsic mechanical modes of the optical elements[18]. sured Q-factors are only limited by temperature depen- Yet, in general, the required high reflectivity of the micro- dent intrinsic dissipation[32, 33], which can be reduced arXiv:0802.4384v2 [quant-ph] 18 Jul 2008 element limits its dimension to wavelength size and thus by low temperature operation. Moreover, the observed sets an upper limit to the mechanical frequencies that geometry dependent loss mediated by normal mode cou- can be achieved. Also, it is exceedingly difficult to at- pling may be of relevance across a wide range of micro- tain high mechanical quality factors while maintaining and nanomechanical oscillators, for which a detailed un- high optical reflectivity as independent control of opti- derstanding of dissipation is lacking. cal and mechanical degrees of freedom is generally not possible. Thus, although remarkably high mechanical Q- factors at low frequencies have been obtained[17], this RESULTS and previous approaches[3, 4, 5, 15, 16, 18] have so far not succeeded in combining mechanical Q-factors com- Starting point of our analysis are toroidal silica parable to those achieved in the field of NEMS and microcavites[34] that intrinsically combine ultra-high-Q MEMS[19, 20, 21, 22, 23] with the best values of op- optical whispering gallery modes with around 20 me- tical finesse[24, 25, 26]. But exactly the combination chanical modes in the 0-100 MHz range[9] which are 2 FIG. 1: Observation of mechanical mode coupling. a Mechanical Q-factors (upper panel) and frequencies (lower panel, where solid lines denote results of an FEM simulation) of the radial breathing mode (inset) for varying relative undercut u = L=R. The Q-factors were found to be remarkably reproducible for six different samples and strongly geometry dependent due to intermode coupling. b Q-factors (upper panel) and frequencies (lower panel) of radial breathing mode and a flexural mode of one toroid reveal an avoided crossing confirming that the dispersion lines of both modes do not cross. The mode patterns (radial and flexural modes) hybridize when approaching the coupling region while the corresponding mode patterns switch dispersion lines during the avoided crossing. A coupled harmonic oscillator model (solid lines: coupled Q-factors and frequencies; dashed lines: bare Q-factors and frequencies) allows an excellent fit to the data. observable in an interferometric readout using the struc- onator (cf. Fig. 1a) is varied. To this end consecutive ture's optical modes (see methods). Recent work has XeF2 etching cycles, undercutting the silica structure but shown that the intrinsic coupling of optical and mechani- leaving the silica itself unaffected, are applied. Thus, the cal modes of toroidal microcavities via radiation-pressure relative undercut u = L=R (R: radius of the cavity, L: can give rise to the effect of dynamical backaction[2], length of the free standing membrane, cf. inset of Fig. which allows realization of narrow bandwidth photonic 1a) can be controlled. The dependence of the measured oscillators[27] or photonic RF down-converters[35] as Q-factors of the radial breathing mode (RBM) on the rel- well as radiation pressure cooling of a mechanical ative undercut is depicted in Fig. 1a. Intuitively, it may oscillator[5]. The radial breathing mode (cf. Fig. 1a) in be expected that higher Q-factors are attained for larger particular exhibits strong optomechanical coupling, low undercut due to a reduced clamping area. Interestingly, effective mass (≈ 10−11 kg) and high frequency enabling however, the Q-factors of six different microresonators of the first demonstration of resolved sideband cooling of similar size show a strongly non-monotonous dependence a micromechanical oscillator[30]. To understand their on the relative undercut which contradicts the simple ex- mechanical Q-factors which have remained completely pectation of smaller losses for smaller clamping areas. unexplored so far, we first present a study of mechanical Moreover, the behaviour is remarkably reproducible for dissipation in microtoroids and subsequently show how the different samples. A plot of the measured frequen- the results from this study can be harnessed to devise cies along with an FEM simulation (see methods) of the structures with unprecedentedly low dissipation in the toroids' radially symmetric mechanical modes depicted in frequency range above 30 MHz. Fig. 1a shows excellent agreement of measured and sim- ulated frequencies. Furthermore, the FEM simulation in- dicates an avoided crossing between different mechanical modes as first predicted by Mindlin in 1951[29]. A mea- Observation of mechanical normal mode coupling surement with a different sample allows directly observ- ing this avoided crossing by employing highly sensitive quantum limited optical displacement sensing (see meth- In order to assess the contribution of clamping losses, ods) which allows to also monitor the flexural modes ex- the diameter of the silicon pillar holding the silica res- 3 (0) (1) angular frequencies Ωr=f = Ωr=f + Ωr=f u and Q-factors (0) (1) Qr=f = Qr=f + Qr=f u on the undercut u, these can be used to asymptotically fit the data (cf. Fig. 1b). Using the values obtained from this first fit, only g remains as free parameter in equation (1). A least square fit to both the measured frequencies and Q-factors yields a coupling rate of g = 14 MHz=2π, much larger than the damping rates Γr=f =2π. The corresponding coupled frequencies Ω±=2π and Q-factors Q± match the data remarkably well (cf. Fig. 1b) confirming that indeed both modes behave as two coupled harmonic oscillators giving rise to an avoided crossing between the RBM and the low-Q flexural mode. Moreover, the data presented here, to FIG. 2: Linear relation between D and measured Q- the authors' best knowledge, manifest the first direct factors. Mechanical Q-factors (dots with error bars, left observation of normal mode coupling of two different axis) and corresponding simulated D values (dashed lines, mechanical modes within a micromechanical resonator. right axis) for the radial breathing and a flexural mode of an on-chip toroid show a linear relation in the parameter range depicted where Q ≈ 3 · D. Inset: Mechanical Q-factors de- pending on the background gas pressure.
Recommended publications
  • Resonance Beyond Frequency-Matching
    Resonance Beyond Frequency-Matching Zhenyu Wang (王振宇)1, Mingzhe Li (李明哲)1,2, & Ruifang Wang (王瑞方)1,2* 1 Department of Physics, Xiamen University, Xiamen 361005, China. 2 Institute of Theoretical Physics and Astrophysics, Xiamen University, Xiamen 361005, China. *Corresponding author. [email protected] Resonance, defined as the oscillation of a system when the temporal frequency of an external stimulus matches a natural frequency of the system, is important in both fundamental physics and applied disciplines. However, the spatial character of oscillation is not considered in the definition of resonance. In this work, we reveal the creation of spatial resonance when the stimulus matches the space pattern of a normal mode in an oscillating system. The complete resonance, which we call multidimensional resonance, is a combination of both the spatial and the conventionally defined (temporal) resonance and can be several orders of magnitude stronger than the temporal resonance alone. We further elucidate that the spin wave produced by multidimensional resonance drives considerably faster reversal of the vortex core in a magnetic nanodisk. Our findings provide insight into the nature of wave dynamics and open the door to novel applications. I. INTRODUCTION Resonance is a universal property of oscillation in both classical and quantum physics[1,2]. Resonance occurs at a wide range of scales, from subatomic particles[2,3] to astronomical objects[4]. A thorough understanding of resonance is therefore crucial for both fundamental research[4-8] and numerous related applications[9-12]. The simplest resonance system is composed of one oscillating element, for instance, a pendulum. Such a simple system features a single inherent resonance frequency.
    [Show full text]
  • Normal Modes of the Earth
    Proceedings of the Second HELAS International Conference IOP Publishing Journal of Physics: Conference Series 118 (2008) 012004 doi:10.1088/1742-6596/118/1/012004 Normal modes of the Earth Jean-Paul Montagner and Genevi`eve Roult Institut de Physique du Globe, UMR/CNRS 7154, 4 Place Jussieu, 75252 Paris, France E-mail: [email protected] Abstract. The free oscillations of the Earth were observed for the first time in the 1960s. They can be divided into spheroidal modes and toroidal modes, which are characterized by three quantum numbers n, l, and m. In a spherically symmetric Earth, the modes are degenerate in m, but the influence of rotation and lateral heterogeneities within the Earth splits the modes and lifts this degeneracy. The occurrence of the Great Sumatra-Andaman earthquake on 24 December 2004 provided unprecedented high-quality seismic data recorded by the broadband stations of the FDSN (Federation of Digital Seismograph Networks). For the first time, it has been possible to observe a very large collection of split modes, not only spheroidal modes but also toroidal modes. 1. Introduction Seismic waves can be generated by different kinds of sources (tectonic, volcanic, oceanic, atmospheric, cryospheric, or human activity). They are recorded by seismometers in a very broad frequency band. Modern broadband seismometers which equip global seismic networks (such as GEOSCOPE or IRIS/GSN) record seismic waves between 0.1 mHz and 10 Hz. Most seismologists use seismic records at frequencies larger than 10 mHz (e.g. [1]). However, the very low frequency range (below 10 mHz) has also been used extensively over the last 40 years and provides unvaluable information on the whole Earth.
    [Show full text]
  • 22.51 Course Notes, Chapter 9: Harmonic Oscillator
    9. Harmonic Oscillator 9.1 Harmonic Oscillator 9.1.1 Classical harmonic oscillator and h.o. model 9.1.2 Oscillator Hamiltonian: Position and momentum operators 9.1.3 Position representation 9.1.4 Heisenberg picture 9.1.5 Schr¨odinger picture 9.2 Uncertainty relationships 9.3 Coherent States 9.3.1 Expansion in terms of number states 9.3.2 Non-Orthogonality 9.3.3 Uncertainty relationships 9.3.4 X-representation 9.4 Phonons 9.4.1 Harmonic oscillator model for a crystal 9.4.2 Phonons as normal modes of the lattice vibration 9.4.3 Thermal energy density and Specific Heat 9.1 Harmonic Oscillator We have considered up to this moment only systems with a finite number of energy levels; we are now going to consider a system with an infinite number of energy levels: the quantum harmonic oscillator (h.o.). The quantum h.o. is a model that describes systems with a characteristic energy spectrum, given by a ladder of evenly spaced energy levels. The energy difference between two consecutive levels is ∆E. The number of levels is infinite, but there must exist a minimum energy, since the energy must always be positive. Given this spectrum, we expect the Hamiltonian will have the form 1 n = n + ~ω n , H | i 2 | i where each level in the ladder is identified by a number n. The name of the model is due to the analogy with characteristics of classical h.o., which we will review first. 9.1.1 Classical harmonic oscillator and h.o.
    [Show full text]
  • Normal Modes (Free Oscillations)
    Introduction to Seismology: Lecture Notes 22 April 2005 SEISMOLOGY: NORMAL MODES (FREE OSCILLATIONS) DECOMPOSING SEISMOLOGY INTO SUBDISCIPLINES Seismology can be decomposed into three representative subdisciplines: body waves, surface waves, and normal modes of free oscillation. Technically, these domains form a continuum, each pertaining to particular frequency bands, spatial scales, etc. In all cases, these representations satisfy the wave equation, but each is subject to different boundary conditions and simplifying assumptions. Each is therefore relevant to particular types of subsurface investigation. Below is a table summarizing the salient characteristics of the three. Boundary Seismic Domains Type Application Data Conditions Body Waves P-SV SH High frequency travel times; waveforms unbounded dispersion; group c(w) & Surface Waves Rayleigh Love Lithosphere phase u(w) velocities interfaces spherical Normal Modes Spheroidal Modes Toroidal Modes Global power spectra earth As the table suggests, the normal modes provide a framework for representing global seismic waves. Typically, these modes of free oscillation are of extremely low frequency and are therefore difficult to observe in seismograms. Only the most energetic earthquakes are capable of generating free oscillations that are readily apparent on most seismograms, and then only if the seismograms extend over several days. NORMAL MODES To understand normal modes, which describe the modes of free oscillation of a sphere, it’s instructive to consider the 1D analog of a vibrating string fixed at both ends as shown in panel Figure 1b. This is useful because the 3D case (Figure 1c), similar to the 1D case, requires that Figure 1 Figure by MIT OCW. 1 Introduction to Seismology: Lecture Notes 22 April 2005 standing waves ‘wrap around’ and meet at a null point.
    [Show full text]
  • Normal Mode Analysis ©David Ronis Mcgill University
    Chemistry 365: Normal Mode Analysis ©David Ronis McGill University 1. Quantum Mechanical Treatment Our starting point is the Schrodinger wav e equation: N −2 ∂2 − h + → → Ψ → → = Ψ → → Σ → U(r1,...,r N ) (r1,...,r N ) E (r1,...,r N ), (1.1) = ∂ 2 i 1 2mi ri where N is the number of atoms in the molecule, mi is the mass of the i’th atom, and → → U(r1,...,r N )isthe effective potential for the nuclear motion, e.g., as is obtained in the Born- Oppenheimer approximation. If the amplitude of the vibrational motion is small, then the vibrational part of the Hamil- tonian associated with Eq. (1.1) can be written as: N −2 ∂2 N ↔ → → ≈− h + + 1 ∆ ∆ Hvib Σ →2 U0 Σ Ki, j: i j,(1.2) i=1 2mi ∂∆ 2 i, j=1 i → ∆ ≡ → − → → where U0 is the minimum value of the potential energy, i ri Ri, Ri is the equilibrium posi- tion of the i’th atom, and 2 ↔ ∂ ≡ U Ki, j → → (1.3) ∂ ∂ → ri r j → r k = Rk is the matrix of (harmonic) force constants. Henceforth, we will shift the zero of energy so as to = make U0 0. Note that in obtaining Eq. (1.2), we have neglected anharmonic (i.e., cubic and higher order) corrections to the vibrational motion. The next and most confusing step is to change to matrix notation. We introduce a column vector containing the displacements as: ∆≡ ∆x ∆y ∆z ∆x ∆y ∆z T [ 1 , 1, 1,..., N , N , N ] ,(1.4) where "T"denotes a matrix transpose.
    [Show full text]
  • Waves & Normal Modes
    Waves & Normal Modes Matt Jarvis February 2, 2016 Contents 1 Oscillations 2 1.0.1 Simple Harmonic Motion - revision . 2 2 Normal Modes 5 2.1 Thecoupledpendulum.............................. 6 2.1.1 TheDecouplingMethod......................... 7 2.1.2 The Matrix Method . 10 2.1.3 Initial conditions and examples . 13 2.1.4 Energy of a coupled pendulum . 15 2.2 Unequal Coupled Pendula . 18 2.3 The Horizontal Spring-Mass system . 22 2.3.1 Decouplingmethod............................ 22 2.3.2 The Matrix Method . 23 2.3.3 Energy of the horizontal spring-mass system . 25 2.3.4 Initial Condition . 26 2.4 Vertical spring-mass system . 26 2.4.1 The matrix method . 27 2.5 Interlude: Solving inhomogeneous 2nd order di↵erential equations . 28 2.6 Horizontal spring-mass system with a driving term . 31 2.7 The Forced Coupled Pendulum with a Damping Factor . 33 3 Normal modes II - towards the continuous limit 39 3.1 N-coupled oscillators . 39 3.1.1 Special cases . 40 3.1.2 General case . 42 3.1.3 N verylarge ............................... 44 3.1.4 Longitudinal Oscillations . 47 4WavesI 48 4.1 Thewaveequation ................................ 48 4.1.1 TheStretchedString........................... 48 4.2 d’Alambert’s solution to the wave equation . 50 4.2.1 Interpretation of d’Alambert’s solution . 51 4.2.2 d’Alambert’s solution with boundary conditions . 52 4.3 Solving the wave equation by separation of variables . 54 4.3.1 Negative C . 55 i 1 4.3.2 Positive C . 56 4.3.3 C=0.................................... 56 4.4 Sinusoidalwaves ................................
    [Show full text]
  • The 1D Schrödinger Equation for a Free Particle
    Lecture 3 Phys 3750 Two Coupled Oscillators / Normal Modes Overview and Motivation: Today we take a small, but significant, step towards wave motion. We will not yet observe waves, but this step is important in its own right. The step is the coupling together of two oscillators via a spring that is attached to both oscillating objects. Key Mathematics: We gain some experience with coupled, linear ordinary differential equations. In particular we find special solutions to these equations, known as normal modes, by solving an eigenvalue problem. I. Two Coupled Oscillators Let's consider the diagram shown below, which is nothing more than 2 copies of an harmonic oscillator, the system that we discussed last time. We assume that both oscillators have the same mass m and spring constant k s . Notice, however, that because there are two oscillators each has it own displacement, either q1 or q2 . ks m m ks q1 q2 q1 = 0 q2 = 0 Based on the discussion last time you should be able to immediately write down the equations of motion (one for each oscillating object) as ~2 q&&1 +ω q1 = 0, and (1a) ~ 2 q&&2 + ω q2 = 0 , (1b) ~ 2 where ω = k s m . As we saw last time, the solution to each of theses equations is harmonic motion at the (angular) frequency ω~ . As should be obvious from the D M Riffe -1- 1/4/2013 Lecture 3 Phys 3750 picture, the motion of each oscillator is independent of the other oscillator. This is also reflected in the equation of motion for each oscillator, which has nothing to do with the other oscillator.
    [Show full text]
  • What Is a Photon? Foundations of Quantum Field Theory
    What is a Photon? Foundations of Quantum Field Theory C. G. Torre June 16, 2018 2 What is a Photon? Foundations of Quantum Field Theory Version 1.0 Copyright c 2018. Charles Torre, Utah State University. PDF created June 16, 2018 Contents 1 Introduction 5 1.1 Why do we need this course? . 5 1.2 Why do we need quantum fields? . 5 1.3 Problems . 6 2 The Harmonic Oscillator 7 2.1 Classical mechanics: Lagrangian, Hamiltonian, and equations of motion . 7 2.2 Classical mechanics: coupled oscillations . 8 2.3 The postulates of quantum mechanics . 10 2.4 The quantum oscillator . 11 2.5 Energy spectrum . 12 2.6 Position, momentum, and their continuous spectra . 15 2.6.1 Position . 15 2.6.2 Momentum . 18 2.6.3 General formalism . 19 2.7 Time evolution . 20 2.8 Coherent States . 23 2.9 Problems . 24 3 Tensor Products and Identical Particles 27 3.1 Definition of the tensor product . 27 3.2 Observables and the tensor product . 31 3.3 Symmetric and antisymmetric tensors. Identical particles. 32 3.4 Symmetrization and anti-symmetrization for any number of particles . 35 3.5 Problems . 36 4 Fock Space 38 4.1 Definitions . 38 4.2 Occupation numbers. Creation and annihilation operators. 40 4.3 Observables. Field operators. 43 4.3.1 1-particle observables . 43 4.3.2 2-particle observables . 46 4.3.3 Field operators and wave functions . 47 4.4 Time evolution of the field operators . 49 4.5 General formalism . 51 3 4 CONTENTS 4.6 Relation to the Hilbert space of quantum normal modes .
    [Show full text]
  • Normal Modes and Quality Factors of Spherical Dielectric Resonators: I – Shielded Dielectric Sphere
    PRAMANA °c Indian Academy of Sciences Vol. 62, No. 6 | journal of June 2004 physics pp. 1255{1271 Normal modes and quality factors of spherical dielectric resonators: I { Shielded dielectric sphere R A YADAV and I D SINGH Spectroscopy Laboratory, Department of Physics, Banaras Hindu University, Varanasi 221 005, India E-mail: [email protected] MS received 16 July 2002; revised 27 October 2003; accepted 1 April 2004 Abstract. Electromagnetic theoretic analysis of shielded homogeneous and isotropic di- electric spheres has been made. Characteristic equations for the TE and TM modes have been derived. Dielectric spheres of radii of the order of ¹m size are found suitable for the optical frequency region whereas for the microwave region radii of the order of mm size are found suitable. Parameters suitable for their application in the optical and microwave frequency ranges have been used to compute the frequencies corresponding to the normal modes for the TE and TM modes. Expressions for the quality factors for realistic res- onators, i.e., for a dielectric sphere with a non-zero conductivity and a metal shield with a ¯nite conductivity have also been derived for the TE and TM modes. Computations of the quality factors have been made for resonators with parameters suitable for the optical and the microwave regions. Keywords. Eigenmodes; spherical resonators; spherical dielectric resonators; quality factors. PACS No. 42.50.Dv 1. Introduction The earliest reported study on spherical resonators seems to be that of Debye [1] wherein he has studied normal modes of a conducting sphere embedded in a perfect dielectric medium.
    [Show full text]
  • Calculation of Molecular Vibrational Normal Modes
    Calculation of Molecular Vibrational Normal Modes Benjamin Rosman 0407237H September 4, 2008 Supervisor Dr Alex Welte Abstract Normal mode analysis provides a vital key to understanding the dynamics of a complicated system. In this case, this is the motion and vibrations of the atoms in a molecule. It is shown in several test cases that the algorithm successfully detects every normal mode of the molecule, as well as all rigid body rotations and translations. The frequencies of these vibrations are also computed to within 6% of the values reported in the literature. 1 Introduction The energy and wave function describing the states of a molecule are computed by solving the time- independent Schrodinger¨ Equation. However, for large molecules this can involve finding the solution to partial differential equations of several hundred variables, which can become completely infeasible. This calculation can be simplified by use of the Born-Oppenheimer approximation, which allows for the electronic and nuclear problems to be solved independently [Griffiths 2005]. Research is currently being conducted into different ways of solving the electronic problem. The work presented in this document serves to complement that, by discussing the development of tools which would allow for solving the nuclear component of the molecular energy problem, assuming that the electronic problem had already been resolved. The output from the electronic problem is a grid of potentials, which is not something that is directly measurable in a molecule. Instead observables such as frequencies are calculated by this software. Section 2 provides a discussion of normal modes and their particular relevance to quantum chemistry, including the method used in their computation.
    [Show full text]
  • 8 Surface Waves and Normal Modes
    8 Surface waves and normal modes Our treatment to this point has been limited to body waves, solutions to the seismic wave equation that exist in whole spaces. However, when free surfaces exist in a medium, other solutions are possible and are given the name surface waves. There are two types of surface waves that propagate along Earth’s surface: Rayleigh waves and Love waves. For laterally homogeneous models, Rayleigh waves are radially polarized (P/SV) and exist at any free surface, whereas Love waves are transversely polarized and require some velocity increase with depth (or a spherical geometry). Surface waves are generally the strongest arrivals recorded at teleseismic distances and they provide some of the best constraints on Earth’s shallow structure and low-frequency source properties. They differ from body waves in many respects – they travel more slowly, their amplitude decay with range is generally much less, and their velocities are strongly frequency dependent. Surface waves from large earthquakes are observable for many hours, during which time they circle the Earth multiple times. Constructive interference among these orbiting surface waves, to- gether with analogous reverberations of body waves, form the normal modes,or free oscillations of the Earth. Surface waves and normal modes are generally ob- served at periods longer than about 10 s, in contrast to the much shorter periods seen in many body wave observations. 8.1 Love waves Love waves are formed through the constructive interference of high-order SH surface multiples (i.e., SSS, SSSS, SSSSS, etc.). Thus, it is possible to model Love waves as a sum of body waves.
    [Show full text]
  • Normal Mode and Surface Wave Observations
    1.04 Theory and Observations: Normal Mode and Surface Wave Observations G Laske, Scripps Institution of Oceanography, University of California San Diego, La Jolla, CA, USA R Widmer-Schnidrig, Black Forest Observatory, Wolfach, Germany ã 2015 Elsevier B.V. All rights reserved. 1.04.1 Introduction 117 1.04.1.1 Free Oscillations 117 1.04.1.2 Surface Waves 118 1.04.1.3 What We See in Seismograms: The Basics 119 1.04.2 Free Oscillations 120 1.04.2.1 Modes of a Spherically Symmetric Earth 120 1.04.2.2 Modes of a Heterogeneous Earth 122 1.04.2.2.1 Mode splitting 122 1.04.2.2.2 Mode coupling 126 1.04.2.3 Measuring Mode Observables 128 1.04.2.3.1 Multiplet stripping and degenerate mode frequencies 130 1.04.2.3.2 Singlet stripping and receiver stripping 132 1.04.2.3.3 Retrieving the splitting matrix with the matrix autoregressive technique 133 1.04.2.3.4 Retrieving the splitting matrix with iterative spectral fitting 133 1.04.2.3.5 Observed mode coupling 135 1.04.2.4 Example of a Mode Application: Inner-Core Rotation 136 1.04.2.5 Example of a Mode Application: Earth’s Hum 139 1.04.3 Surface Waves 141 1.04.3.1 Standing Waves and Traveling Waves 141 1.04.3.2 The Measurement of Fundamental Mode Dispersion 144 1.04.3.2.1 Group velocity 148 1.04.3.2.2 The study of ambient noise 149 1.04.3.2.3 Phase velocity 149 1.04.3.2.4 Time-variable filtering 151 1.04.3.3 Other Surface Wave Observables 152 1.04.3.4 Higher Modes 154 1.04.3.4.1 Higher-mode dispersion and waveform modeling 154 1.04.3.4.2 Love waves and overtones 157 1.04.3.5 Surface Waves and Structure at Depth 157 1.04.3.6 The Validity of the Great-Circle Approximation and Data Accuracy 160 1.04.4 Concluding Remarks 160 Acknowledgments 161 References 161 1.04.1 Introduction Valdivia, Chile, earthquake.
    [Show full text]