Quantum-optical spectroscopy

Last updated

Quantum-optical spectroscopy [1] [2] is a quantum-optical generalization of laser spectroscopy where matter is excited and probed with a sequence of laser pulses.

Contents

Classically, such pulses are defined by their spectral and temporal shape as well as phase and amplitude of the electromagnetic field. Besides these properties of light, the phase-amplitude aspects have intrinsic quantum fluctuations that are of central interest in quantum optics. In ordinary laser spectroscopy, [3] [4] [5] one utilizes only the classical aspects of laser pulses propagating through matter such as atoms or semiconductors. In quantum-optical spectroscopy, one additionally utilizes the quantum-optical fluctuations of light to enhance the spectroscopic capabilities by directly shaping and/or detecting the quantum fluctuations of light. Quantum-optical spectroscopy has applications in controlling and characterizing quantum dynamics of many-body states because one can directly access a large set of many-body states, [6] [7] which is not possible in classical spectroscopy .

Quantum-optical state injection

A generic electromagnetic field can always be expressed in terms of a mode expansion where individual components form a complete set of modes. Such modes can be constructed with different methods and they can, e.g., be energy eigenstate, generic spatial modes, or temporal modes. Once these light mode are chosen, their effect on the quantized electromagnetic field can be described by Boson creation and annihilation operators and for photons, respectively. [8] The quantum fluctuations of the light field can be uniquely defined [9] by the photon correlations that contain the pure -particle correlations as defined with the cluster-expansion approach. Using the same second-quantization formalism for the matter being studied, typical electronic excitations in matter can be described by Fermion operators for electronic excitations and holes, i.e.~electronic vacancies left behind to the many-body ground state. [10] The corresponding electron–hole excitations can be described by operators and that create and annihilate an electron–hole pair, respectively.

In several relevant cases, the light–matter interaction can be described using the dipole interaction [7]

where the summation is implicitly taken over all possibilities to create an electron–hole pair (the part) via a photon absorption (the part); the Hamiltonian also contains the Hermitian conjugate (abbreviated as h.c.) of the terms that are explicitly written. The coupling strength between light and matter is defined by .

When the electron–hole pairs are excited resonantly with a single-mode light , the photon correlations are directly injected into the many-body correlations. More specifically, the fundamental form of the light–matter interaction inevitably leads to a correlation-transfer relation [1] [7]

between photons and electron–hole excitations. Strictly speaking, this relation is valid before the onset of scattering induced by the Coulomb and phonon interactions in the solid. Therefore, it is desirable to use laser pulses that are faster than the dominant scattering processes. This regime is relatively easy to realize in present-day laser spectroscopy because lasers can already output femtosecond, or even attosecond, pulses with a high precision in controllability.

Realization

Physically, the correlation-transfer relation means that one can directly inject desired many-body states simply by adjusting the quantum fluctuations of the light pulse, as long as the light pulse is short enough. This opens a new possibility for studying properties of distinct many-body states, once the quantum-optical spectroscopy is realized through controlling the quantum fluctuations of light sources. For example, a coherent-state laser is described entirely by its single-particle expectation value . Therefore, such excitation directly injects property that is polarization related to electron–hole transitions. To directly excite bound electron–hole pairs, i.e., excitons, described by a two-particle correlation , or a biexciton transition , one needs to have a source with or photon correlations, respectively.

To realize quantum-optical spectroscopy, high-intensity light sources with freely adjustable quantum statistics are needed which are currently not available. However, one can apply projective methods [6] [11] [12] to access the quantum–optical response of matter from a set of classical measurements. Especially, the method presented by Kira, M. et al [6] is robust in projecting quantum-optical responses of genuine many-body systems. This work has shown that one can indeed reveal and access many–body properties that remain hidden in classical spectroscopy. Therefore, the quantum-optical spectroscopy is ideally suited for characterizing and controlling complicated many-body states in several different systems, ranging from molecules to semiconductors.

Relation to semiconductor quantum optics

Quantum-optical spectroscopy is an important approach in general semiconductor quantum optics. The capability to discriminate and control many-body states is certainly interesting in extended semiconductors such as quantum wells because a typical classical excitation indiscriminately detects contributions from multiple many-body configurations; With quantum-optical spectroscopy one can access and control a desired many-body state within an extended semiconductor. [7] At the same time, the ideas of quantum-optical spectroscopy can also be useful when studying simpler systems such as quantum dots.

Quantum dots are a semiconductor equivalent to simple atomic systems where most of the first quantum-optical demonstrations have been measured. [8] Since quantum dots are man-made, one can possibly customize them to produce new quantum-optical components for information technology. For example, in quantum-information science, one is often interested to have light sources that can output photons on demand or entangled photon pairs at specific frequencies. Such sources have already been demonstrated with quantum dots by controlling their photon emission with various schemes. [13] [14] [15] In the same way, quantum-dot lasers may exhibit unusual changes in the conditional probability [16] to emit a photon when already one photon is emitted; this effect can be measured in the so-called g2 correlation. One interesting possibility for quantum-optical spectroscopy is to pump quantum dots with quantum light to control their light emission more precisely. [17]

Quantum-dot microcavity investigations have progressed rapidly ever since the experimental demonstration [18] [19] of vacuum Rabi splitting between a single dot and a cavity resonance. This regime can be understood on the basis of the Jaynes–Cummings model while the semiconductor aspects provide many new physical effects [20] [21] due to the electronic coupling with the lattice vibrations.

Nevertheless, the quantum Rabi splitting—stemming directly from the quantized light levels—remained elusive because many experiments were monitoring only the intensity of photoluminescence. Following the ideology of quantum-optical spectroscopy, Ref. [22] predicted that quantum-Rabi splitting could be resolved in photon-correlation measurement even when it becomes smeared out in photoluminescence spectrum. This was experimentally demonstrated [23] by measuring the so-called g2 correlations that quantify how regularly the photons are emitted by the quantum dot inside a microcavity.

See also

Related Research Articles

Photoluminescence Light emission from substances after they absorb photons

Photoluminescence is light emission from any form of matter after the absorption of photons. It is one of many forms of luminescence and is initiated by photoexcitation, hence the prefix photo-. Following excitation, various relaxation processes typically occur in which other photons are re-radiated. Time periods between absorption and emission may vary: ranging from short femtosecond-regime for emission involving free-carrier plasma in inorganic semiconductors up to milliseconds for phosphoresence processes in molecular systems; and under special circumstances delay of emission may even span to minutes or hours.

In physics, specifically in quantum mechanics, a coherent state is the specific quantum state of the quantum harmonic oscillator, often described as a state which has dynamics most closely resembling the oscillatory behavior of a classical harmonic oscillator. It was the first example of quantum dynamics when Erwin Schrödinger derived it in 1926, while searching for solutions of the Schrödinger equation that satisfy the correspondence principle. The quantum harmonic oscillator arise in the quantum theory of a wide range of physical systems. For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well. The coherent state describes a state in a system for which the ground-state wavepacket is displaced from the origin of the system. This state can be related to classical solutions by a particle oscillating with an amplitude equivalent to the displacement.

Squeezed coherent state Type of quantum state

In physics, a squeezed coherent state is a quantum state that is usually described by two non-commuting observables having continuous spectra of eigenvalues. Examples are position and momentum of a particle, and the (dimension-less) electric field in the amplitude and in the mode of a light wave. The product of the standard deviations of two such operators obeys the uncertainty principle:

Electromagnetically induced transparency

Electromagnetically induced transparency (EIT) is a coherent optical nonlinearity which renders a medium transparent within a narrow spectral range around an absorption line. Extreme dispersion is also created within this transparency "window" which leads to "slow light", described below. It is in essence a quantum interference effect that permits the propagation of light through an otherwise opaque atomic medium.

Coherent control is a quantum mechanics-based method for controlling dynamical processes by light. The basic principle is to control quantum interference phenomena, typically by shaping the phase of laser pulses. The basic ideas have proliferated, finding vast application in spectroscopy mass spectra, quantum information processing, laser cooling, ultracold physics and more.

Quantum noise is noise arising from the indeterminate state of matter in accordance with fundamental principles of quantum mechanics, specifically the uncertainty principle and via zero-point energy fluctuations. Quantum noise is due to the apparently discrete nature of the small quantum constituents such as electrons, as well as the discrete nature of quantum effects, such as photocurrents.

Jaynes–Cummings model Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

A NOON state is a quantum-mechanical many-body entangled state:

In physics, a quantum amplifier is an amplifier that uses quantum mechanical methods to amplify a signal; examples include the active elements of lasers and optical amplifiers.

Resonance fluorescence is the process in which a two-level atom system interacts with the quantum electromagnetic field if the field is driven at a frequency near to the natural frequency of the atom.

The Hong–Ou–Mandel effect is a two-photon interference effect in quantum optics that was demonstrated in 1987 by three physicists from the University of Rochester: Chung Ki Hong (홍정기), Zheyu Ou (区泽宇), and Leonard Mandel. The effect occurs when two identical single-photon waves enter a 1:1 beam splitter, one in each input port. When the temporal overlap of the photons on the beam splitter is perfect, the two photons will always exit the beam splitter together in the same output mode, meaning that there is zero chance that they will exit separately with one photon in each of the two outputs giving a coincidence event. The photons have a 50:50 chance of exiting (together) in either output mode. If they become more distinguishable, the probability of them each going to a different detector will increase. In this way, the interferometer coincidence signal can accurately measure bandwidth, path lengths, and timing. Since this effect relies on the existence of photons and the second quantization it can not be fully explained by classical optics.

In his historic paper entitled "The Quantum Theory of Optical Coherence," Roy J. Glauber set a solid foundation for the quantum electronics/quantum optics enterprise. The experimental development of the optical maser and later laser at that time had made the classical concept of optical coherence inadequate. Glauber started from the quantum theory of light detection by considering the process of photoionization in which a photodetector is triggered by an ionizing absorption of a photon. In the quantum theory of radiation, the electric field operator in the Coulomb gauge may be written as the sum of positive and negative frequency parts

The semiconductor Bloch equations describe the optical response of semiconductors excited by coherent classical light sources, such as lasers. They are based on a full quantum theory, and form a closed set of integro-differential equations for the quantum dynamics of microscopic polarization and charge carrier distribution. The SBEs are named after the structural analogy to the optical Bloch equations that describe the excitation dynamics in a two-level atom interacting with a classical electromagnetic field. As the major complication beyond the atomic approach, the SBEs must address the many-body interactions resulting from Coulomb force among charges and the coupling among lattice vibrations and electrons.

The semiconductor luminescence equations (SLEs) describe luminescence of semiconductors resulting from spontaneous recombination of electronic excitations, producing a flux of spontaneously emitted light. This description established the first step toward semiconductor quantum optics because the SLEs simultaneously includes the quantized light–matter interaction and the Coulomb-interaction coupling among electronic excitations within a semiconductor. The SLEs are one of the most accurate methods to describe light emission in semiconductors and they are suited for a systematic modeling of semiconductor emission ranging from excitonic luminescence to lasers.

The interaction of matter with light, i.e., electromagnetic fields, is able to generate a coherent superposition of excited quantum states in the material. Coherent denotes the fact that the material excitations have a well defined phase relation which originates from the phase of the incident electromagnetic wave. Macroscopically, the superposition state of the material results in an optical polarization, i.e., a rapidly oscillating dipole density. The optical polarization is a genuine non-equilibrium quantity that decays to zero when the excited system relaxes to its equilibrium state after the electromagnetic pulse is switched off. Due to this decay which is called dephasing, coherent effects are observable only for a certain temporal duration after pulsed photoexcitation. Various materials such as atoms, molecules, metals, insulators, semiconductors are studied using coherent optical spectroscopy and such experiments and their theoretical analysis has revealed a wealth of insights on the involved matter states and their dynamical evolution.

The cluster-expansion approach is a technique in quantum mechanics that systematically truncates the BBGKY hierarchy problem that arises when quantum dynamics of interacting systems is solved. This method is well suited for producing a closed set of numerically computable equations that can be applied to analyze a great variety of many-body and/or quantum-optical problems. For example, it is widely applied in semiconductor quantum optics and it can be applied to generalize the semiconductor Bloch equations and semiconductor luminescence equations.

Stephan W. Koch is a German theoretical physicist. He is professor at the University of Marburg and works on condensed-matter theory, many-body effects, and laser theory. He is best known for his seminal contributions to the optical and electronic properties of semiconductors, semiconductor quantum optics, and semiconductor laser designs. Major portion of his research work has focused on the quantum physics and application potential of semiconductor nanostructures. Besides gaining fundamental insights to the many-body quantum theory, his work has provided new possibilities to develop, e.g., laser technology, based on accurate computer simulations. His objective has been to self-consistently include all relevant many-body effects in order to eliminate phenomenological approximations that compromise predictability of effects and quantum-device designs.

Semiconductor laser theory Theory of laser diodes

Semiconductor lasers or laser diodes play an important part in our everyday lives by providing cheap and compact-size lasers. They consist of complex multi-layer structures requiring nanometer scale accuracy and an elaborate design. Their theoretical description is important not only from a fundamental point of view, but also in order to generate new and improved designs. It is common to all systems that the laser is an inverted carrier density system. The carrier inversion results in an electromagnetic polarization which drives an electric field . In most cases, the electric field is confined in a resonator, the properties of which are also important factors for laser performance.

Terahertz spectroscopy detects and controls properties of matter with electromagnetic fields that are in the frequency range between a few hundred gigahertz and several terahertz. In many-body systems, several of the relevant states have an energy difference that matches with the energy of a THz photon. Therefore, THz spectroscopy provides a particularly powerful method in resolving and controlling individual transitions between different many-body states. By doing this, one gains new insights about many-body quantum kinetics and how that can be utilized in developing new technologies that are optimized up to the elementary quantum level.

The Mandel Q parameter measures the departure of the occupation number distribution from Poissonian statistics. It was introduced in quantum optics by Leonard Mandel. It is a convenient way to characterize non-classical states with negative values indicating a sub-Poissonian statistics, which have no classical analog. It is defined as the normalized variance of the boson distribution:

References

  1. 1 2 Kira, M.; Koch, S. (2006). "Quantum-optical spectroscopy of semiconductors". Physical Review A73 (1). doi:10.1103/PhysRevA.73.013813. ISSN   1050-2947.
  2. Koch, S. W.; Kira, M.; Khitrova, G.; Gibbs, H. M. (2006). "Semiconductor excitons in new light". Nature Materials5 (7): 523–531. doi:10.1038/nmat1658. ISSN   1476-1122.
  3. Stenholm, S. (2005). Foundations of laser spectroscopy. Dover Pubn. Inc. ISBN   978-0486444987.
  4. Demtröder, W. (2008). Laser Spectroscopy: Vol. 1: Basic Principles. Springer. ISBN   978-3540734154.
  5. Demtröder, W. (2008). Laser Spectroscopy: Vol. 2: Experimental Techniques. Springer. ISBN   978-3540749523.
  6. 1 2 3 Kira, M.; Koch, S. W.; Smith, R. P.; Hunter, A. E.; Cundiff, S. T. (2011). "Quantum spectroscopy with Schrödinger-cat states". Nature Physics7 (10): 799–804. doi:10.1038/nphys2091. ISSN   1745-2473.
  7. 1 2 3 4 Kira, M.; Koch, S. W. (2011). Semiconductor Quantum Optics. Cambridge University Press. ISBN   978-0521875097.
  8. 1 2 Walls, D. F.; Milburn, G. J. (2008). Quantum Optics. Springer. ISBN   978-3540285731.
  9. Kira, M.; Koch, S. (2008). "Cluster-expansion representation in quantum optics". Physical Review A 78 (2). doi:10.1103/PhysRevA.78.022102. ISSN   1050-2947.
  10. Ashcroft, N. W.; Mermin, N. D. (1976). Solid state physics. Saunders College. ISBN   978-0030839931.
  11. Sudarshan, E. (1963). "Equivalence of Semiclassical and Quantum Mechanical Descriptions of Statistical Light Beams". Physical Review Letters10 (7): 277–279. doi:10.1103/PhysRevLett.10.277. ISSN   0031-9007.
  12. Lobino, M.; Korystov, D.; Kupchak, C.; Figueroa, E.; Sanders, B. C.; Lvovsky, A. I. (2008). "Complete Characterization of Quantum-Optical Processes". Science322 (5901): 563–566. doi:10.1126/science.1162086. ISSN   0036-8075.
  13. Michler, P. (2000). "A Quantum Dot Single-Photon Turnstile Device". Science290 (5500): 2282–2285. doi:10.1126/science.290.5500.2282. ISSN   0036-8075.
  14. Benson, Oliver; Santori, Charles; Pelton, Matthew; Yamamoto, Yoshihisa (2000). "Regulated and Entangled Photons from a Single Quantum Dot". Physical Review Letters84 (11): 2513–2516. doi:10.1103/PhysRevLett.84.2513. ISSN   0031-9007.
  15. Stevenson, R. M.; Young, R. J.; Atkinson, P.; Cooper, K.; Ritchie, D. A.; Shields, A. J. (2006). "A semiconductor source of triggered entangled photon pairs". Nature439 (7073): 179–182. doi:10.1038/nature04446. ISSN   0028-0836.
  16. Ulrich, S. M.; Gies, C.; Ates, S.; Wiersig, J.; Reitzenstein, S.; Hofmann, C.; Löffler, A.; Forchel, A.; Jahnke, F.; Michler, P. (2007). "Photon Statistics of Semiconductor Microcavity Lasers". Physical Review Letters98 (4). doi:10.1103/PhysRevLett.98.043906. ISSN   0031-9007.
  17. Aßmann, Marc; Bayer, Manfred (2011). "Nonlinearity sensing via photon-statistics excitation spectroscopy". Physical Review A84 (5). doi:10.1103/PhysRevA.84.053806. ISSN   1050-2947.
  18. Reithmaier, J. P.; Sęk, G.; Löffler, A.; Hofmann, C.; Kuhn, S.; Reitzenstein, S.; Keldysh, L. V.; Kulakovskii, V. D.; Reinecke, T. L.; Forchel, A. (2004). "Strong coupling in a single quantum dot–semiconductor microcavity system". Nature432 (7014): 197–200. doi:10.1038/nature02969. ISSN   0028-0836.
  19. Yoshie, T.; Scherer, A.; Hendrickson, J.; Khitrova, G.; Gibbs, H. M.; Rupper, G.; Ell, C.; Shchekin, O. B. et al. (2004). "Vacuum Rabi splitting with a single quantum dot in a photonic crystal nanocavity". Nature432 (7014): 200–203. doi:10.1038/nature03119. ISSN   0028-0836.
  20. Förstner, J.; Weber, C.; Danckwerts, J.; Knorr, A. (2003). "Phonon-Assisted Damping of Rabi Oscillations in Semiconductor Quantum Dots". Physical Review Letters91 (12). doi:10.1103/PhysRevLett.91.127401. ISSN   0031-9007.
  21. Carmele, Alexander; Richter, Marten; Chow, Weng W.; Knorr, Andreas (2010). "Antibunching of Thermal Radiation by a Room-Temperature Phonon Bath: A Numerically Solvable Model for a Strongly Interacting Light-Matter-Reservoir System". Physical Review Letters104 (15). doi:PhysRevLett.104.156801. ISSN   0031-9007.
  22. Schneebeli, L.; Kira, M.; Koch, S. (2008). "Characterization of Strong Light-Matter Coupling in Semiconductor Quantum-Dot Microcavities via Photon-Statistics Spectroscopy". Physical Review Letters101 (9). doi:10.1103/PhysRevLett.101.097401. ISSN   0031-9007}.
  23. Reinhard, Andreas; Volz, Thomas; Winger, Martin; Badolato, Antonio; Hennessy, Kevin J.; Hu, Evelyn L.; Imamoğlu, Ataç (2011). "Strongly correlated photons on a chip". Nature Photonics6 (2): 93–96. doi:10.1038/nphoton.2011.321. ISSN   1749-4885.

Further reading