Two-photon absorption

Last updated
Schematic of energy levels involved in two photons absorption Two photons absorption energy scheme.png
Schematic of energy levels involved in two photons absorption

In atomic physics, two-photon absorption (TPA or 2PA), also called two-photon excitation or non-linear absorption, is the simultaneous absorption of two photons of identical or different frequencies in order to excite an atom or a molecule from one state (usually the ground state) to a higher energy, most commonly an excited electronic state. Absorption of two photons with different frequencies is called non-degenerate two-photon absorption. Since TPA depends on the simultaneous absorption of two photons, the probability of TPA is proportional to the square of the light intensity; thus it is a nonlinear optical process. [1] The energy difference between the involved lower and upper states of the molecule is equal or smaller than the sum of the photon energies of the two photons absorbed. Two-photon absorption is a third-order process, with absorption cross section typically several orders of magnitude smaller than one-photon absorption cross section.

Contents

Two-photon excitation of a fluorophore (a fluorescent molecule) leads to two-photon-excited fluorescence where the excited state produced by TPA decays by spontaneous emission of a photon to a lower energy state.

Background

The phenomenon was originally predicted by Maria Goeppert-Mayer in 1931 in her doctoral dissertation. [2] Thirty years later, the invention of the laser permitted the first experimental verification of TPA when two-photon-excited fluorescence was detected in a europium-doped crystal. [3] Soon afterwards, the effect was observed in cesium vapor and then in CdS, a semiconductor. [4] [5]

Schematic of energy levels involved in two photons excited fluorescence. First there is a two-photons absorption, followed by one non-radiative deexcitation and a fluorescence emission. The electron returns at ground state by another non-radiative deexcitation. The created pulsation
o
2
{\displaystyle \omega _{2}}
is thus smaller than twice the excited pulsation
o
1
{\displaystyle \omega _{1}} Two photons excited fluorescence energy levels.png
Schematic of energy levels involved in two photons excited fluorescence. First there is a two-photons absorption, followed by one non-radiative deexcitation and a fluorescence emission. The electron returns at ground state by another non-radiative deexcitation. The created pulsation is thus smaller than twice the excited pulsation

TPA is a nonlinear optical process. In particular, the imaginary part of the third-order nonlinear susceptibility is related to the extent of TPA in a given molecule. The selection rules for TPA are therefore different from one-photon absorption (OPA), which is dependent on the first-order susceptibility. The relationship between the selection rules for one- and two-photon absorption is analogous to those of Raman and IR spectroscopies. For example, in a centrosymmetric molecule, one- and two-photon allowed transitions are mutually exclusive, an optical transition allowed in one of the spectroscopies is forbidden in the other. However, for non-centrosymmetric molecules there is no formal mutual exclusion between the selection rules for OPA and TPA. In quantum mechanical terms, this difference results from the fact that the quantum states of such molecules have either + or - inversion symmetry, usually labelled by g (for +) and u (for -). One photon transitions are only allowed between states that differ in the inversion symmetry, i.e. g <-> u, while two photon transitions are only allowed between states that have the same inversion symmetry, i.e. g <-> g and u <-> u.

The relation between the number of photons - or, equivalently, order of the electronic transitions - involved in a TPA process (two) and the order of the corresponding nonlinear susceptibility (three) may be understood using the optical theorem. This theorem relates the imaginary part of an all-optical process of a given perturbation order with a process involving charge carriers with half the perturbation order, i.e. . [6] To apply this theorem it is important to consider that the order in perturbation theory to calculate the probability amplitude of an all-optical process is . Since in the case of TPA there are electronic transitions of the second order involved (), it results from the optical theorem that the order of the nonlinear susceptibility is , i.e. it is a process.

Phenomenologically, TPA can be thought of as the third term in a conventional anharmonic oscillator model for depicting vibrational behavior of molecules. Another view is to think of light as photons. In nonresonant TPA, neither photon is at resonance with the system energy gap, and two photons combine to bridge the energy gap larger than the energies of each photon individually. If there were an intermediate electronic state in the gap, this could happen via two separate one-photon transitions in a process described as "resonant TPA", "sequential TPA", or "1+1 absorption" where the absorption alone is a first order process and the generated fluorescence will rise as the square of the incoming intensity. In nonresonant TPA the transition occurs without the presence of the intermediate state. This can be viewed as being due to a "virtual" state created by the interaction of the photons with the molecule. The virtual state argument is quite orthogonal to the anharmonic oscillator argument. It states for example that in a semiconductor, absorption at high energies is impossible if two photons cannot bridge the band gap. So, many materials can be used for the Kerr effect that do not show any absorption and thus have a high damage threshold.

The "nonlinear" in the description of this process means that the strength of the interaction increases faster than linearly with the electric field of the light. In fact, under ideal conditions the rate of TPA is proportional to the square of the field intensity. This dependence can be derived quantum mechanically, but is intuitively obvious when one considers that it requires two photons to coincide in time and space. This requirement for high light intensity means that lasers are required to study TPA phenomena. Further, in order to understand the TPA spectrum, monochromatic light is also desired in order to measure the TPA cross section at different wavelengths. Hence, tunable pulsed lasers (such as frequency-doubled Nd:YAG-pumped OPOs and OPAs) are the choice of excitation.

Measurements

Two-photon absorption can be measured by several techniques. Some of them are two-photon excited fluorescence (TPEF), [7] z-scan, self-diffraction [8] or nonlinear transmission (NLT). Pulsed lasers are most often used because TPA is a third-order nonlinear optical process, [9] and therefore is most efficient at very high intensities.

Absorption rate

Beer's law describes the decay in intensity due to one-photon absorption:

where are the distance that light travelled through a sample, is the light intensity after travelling a distance , is the light intensity where the light enters the sample and is the one-photon absorption coefficient of the sample. In two-photon absorption, for an incident plane wave of radiation, the light intensity versus distance changes to

for TPA with light intensity as a function of path length or cross section as a function of concentration and the initial light intensity . The absorption coefficient now becomes the TPA coefficient. (Note that there is some confusion over the term in nonlinear optics, since it is sometimes used to describe the second-order polarizability, and occasionally for the molecular two-photon cross-section. More often however, it is used to describe the bulk 2-photon optical density of a sample. The letter or is more often used to denote the molecular two-photon cross-section.)

Two-photon excited fluorescence

Relation between the two-photon excited fluorescence and the total number of absorbed photons per unit time is given by

where and are the fluorescence quantum efficiency of the fluorophore and the fluorescence collection efficiency of the measurement system, respectively. [10] In a particular measurement, is a function of fluorophore concentration , illuminated sample volume , incident light intensity , and two-photon absorption cross-section :

Notice that the is proportional to the square of the incident light as expected for TPA.

Units of cross-section

The molecular two-photon absorption cross-section is usually quoted in the units of Goeppert-Mayer (GM) (after its discoverer, Nobel laureate Maria Goeppert-Mayer), where 1 GM is 10−50 cm4 s photon−1. [11] Considering the reason for these units, one can see that it results from the product of two areas (one for each photon, each in cm2) and a time (within which the two photons must arrive to be able to act together). The large scaling factor is introduced in order that 2-photon absorption cross-sections of common dyes will have convenient values.

Development of the field and potential applications

Until the early 1980s, TPA was used as a spectroscopic tool. Scientists compared the OPA and TPA spectra of different organic molecules and obtained several fundamental structure property relationships. However, in late 1980s, applications started to be developed. Peter Rentzepis suggested applications in 3D optical data storage. Watt Webb suggested microscopy and imaging. Other applications such as 3D microfabrication, optical logic, autocorrelation, pulse reshaping and optical power limiting were also demonstrated. [12]

3D imaging of semiconductors

It was demonstrated that by using 2-photon absorption charge carriers can be generated spatially confined in a semiconductor device. This can be used to investigate the charge transport properties of such device. [13]

Microfabrication and lithography

In 1992, with the use of higher laser powers (35 mW) and more sensitive resins/resists, TPA found its way into lithography. [14] One of the most distinguishing features of TPA is that the rate of absorption of light by a molecule depends on the square of the light's intensity. This is different from OPA, where the rate of absorption is linear with respect to input intensity. As a result of this dependence, if material is cut with a high power laser beam, the rate of material removal decreases very sharply from the center of the beam to its periphery. Because of this, the "pit" created is sharper and better resolved than if the same size pit were created using normal absorption.

3D photopolymerization

In 1997, Maruo et al. developed the first application of TPA in 3D microfabrication. [15] In 3D microfabrication, a block of gel containing monomers and a 2-photon active photoinitiator is prepared as a raw material. Application of a focused laser to the block results in polymerization only at the focal spot of the laser, where the intensity of the absorbed light is highest. The shape of an object can therefore be traced out by the laser, and then the excess gel can be washed away to leave the traced solid. Photopolymerization for 3D microfabrication is used in a wide variety of applications, including microoptics [16] , microfluids [17] , biomedical implants [18] , 3D scaffolds for cell cultures [19] and tissue engineering [20] .

Imaging

The human body is not transparent to visible wavelengths. Hence, one photon imaging using fluorescent dyes is not very efficient. If the same dye had good two-photon absorption, then the corresponding excitation would occur at approximately two times the wavelength at which one-photon excitation would have occurred. As a result, it is possible to use excitation in the far infrared region where the human body shows good transparency.

It is sometimes said, incorrectly, that Rayleigh scattering is relevant to imaging techniques such as two-photon. According to Rayleigh's scattering law, the amount of scattering is proportional to , where is the wavelength. As a result, if the wavelength is increased by a factor of 2, the Rayleigh scattering is reduced by a factor of 16. However, Rayleigh scattering only takes place when scattering particles are much smaller than the wavelength of light (the sky is blue because air molecules scatter blue light much more than red light). When particles are larger, scattering increases approximately linearly with wavelength: hence clouds are white since they contain water droplets. This form of scatter is known as Mie scattering and is what occurs in biological tissues. So, although longer wavelengths do scatter less in biological tissues, the difference is not as dramatic as Rayleigh's law would predict.

Optical power limiting

Another area of research is optical power limiting. In a material with a strong nonlinear effect, the absorption of light increases with intensity such that beyond a certain input intensity the output intensity approaches a constant value. Such a material can be used to limit the amount of optical power entering a system. This can be used to protect expensive or sensitive equipment such as sensors, can be used in protective goggles, or can be used to control noise in laser beams.

Photodynamic therapy

Photodynamic therapy (PDT) is a method for treating cancer. In this technique, an organic molecule with a good triplet quantum yield is excited so that the triplet state of this molecule interacts with oxygen. The ground state of oxygen has triplet character. This leads to triplet-triplet annihilation, which gives rise to singlet oxygen, which in turn attacks cancerous cells. However, using TPA materials, the window for excitation can be extended into the infrared region, thereby making the process more viable to be used on the human body.

Two-photon pharmacology

Photoisomerization of azobenzene-based pharmacological ligands by 2-photon absorption has been described for use in photopharmacology. [21] [22] [23] [24] [25] [26] It allows controlling the activity of endogenous proteins in intact tissue with pharmacological selectivity in three dimensions. It can be used to study neural circuits and to develop drug-based non invasive phototherapies.

Optical data storage

The ability of two-photon excitation to address molecules deep within a sample without affecting other areas makes it possible to store and retrieve information in the volume of a substance rather than only on a surface as is done on the DVD. Therefore, 3D optical data storage has the possibility to provide media that contain terabyte-level data capacities on a single disc.

Compounds

To some extent, linear and 2-photon absorption strengths are linked. Therefore, the first compounds to be studied (and many that are still studied and used in e.g. 2-photon microscopy) were standard dyes. In particular, laser dyes were used, since these have good photostability characteristics. However, these dyes tend to have 2-photon cross-sections of the order of 0.1–10 GM, much less than is required to allow simple experiments.

It was not until the 1990s that rational design principles for the construction of two-photon-absorbing molecules began to be developed, in response to a need from imaging and data storage technologies, and aided by the rapid increases in computer power that allowed quantum calculations to be made. The accurate quantum mechanical analysis of two-photon absorbance is orders of magnitude more computationally intensive than that of one-photon absorbance, requiring highly correlated calculations at very high levels of theory.

The most important features of strongly TPA molecules were found to be a long conjugation system (analogous to a large antenna) and substitution by strong donor and acceptor groups (which can be thought of as inducing nonlinearity in the system and increasing the potential for charge-transfer). Therefore, many push-pull olefins exhibit high TPA transitions, up to several thousand GM. [27] It is also found that compounds with a real intermediate energy level close to the "virtual" energy level can have large 2-photon cross-sections as a result of resonance enhancement. There are several databases of two-photon absorption spectra available online. [28] [29]

Compounds with interesting TPA properties also include various porphyrin derivatives, conjugated polymers and even dendrimers. In one study [30] a diradical resonance contribution for the compound depicted below was also linked to efficient TPA. The TPA wavelength for this compound is 1425 nanometer with observed TPA cross section of 424 GM.

DiradicalApplicationinTPA.png

Coefficients

The two-photon absorption coefficient is defined by the relation [31]

so that

Where is the two-photon absorption coefficient, is the absorption coefficient, is the transition rate for TPA per unit volume, is the irradiance, ħ is the reduced Planck constant, is the photon frequency and the thickness of the slice is . is the number density of molecules per cm3, is the photon energy (J), is the two-photon absorption cross section (cm4s/molecule).

The SI units of the beta coefficient are m/W. If (m/W) is multiplied by 10−9 it can be converted to the CGS system (cal/cm s/erg). [32]

Due to different laser pulses the TPA coefficients reported has differed as much as a factor 3. With the transition towards shorter laser pulses, from picosecond to subpicosecond durations, noticeably reduced TPA coefficient have been obtained. [33]

In water

Laser induced TPA in water was discovered in 1980. [34]

Water absorbs UV radiation near 125 nm exiting the 3a1 orbital leading to dissociation into OH and H+. Through TPA this dissociation can be achieved by two photons near 266 nm. [35] Since water and heavy water have different vibration frequencies and inertia they also need different photon energies to achieve dissociation and have different absorption coefficients for a given photon wavelength. A study from Jan 2002 used a femtosecond laser tuned to 0.22 Picoseconds found the coefficient of D2O to be 42±5 10−11(cm/W) whereas H2O was 49±5 10−11(cm/W). [33]

TPA Coefficients for Water [33]
λ (nm)pulse duration τ (ps) (cm/W)
315294
300294.5
289296
282297
2820.1819
2662910
2640.2249±5
2161520
2132632

Two-photon emission

The opposite process of TPA is two-photon emission (TPE), which is a single electron transition accompanied by the emission of a photon pair. The energy of each individual photon of the pair is not determined, while the pair as a whole conserves the transition energy. The spectrum of TPE is therefore very broad and continuous. [36] TPE is important for applications in astrophysics, contributing to the continuum radiation from planetary nebulae (theoretically predicted for them in [37] and observed in [38] ). TPE in condensed matter and specifically in semiconductors was only first observed in 2008, [39] with emission rates nearly 5 orders of magnitude weaker than one-photon spontaneous emission, with potential applications in quantum information.

See also

Related Research Articles

<span class="mw-page-title-main">Nonlinear optics</span> Branch of physics

Nonlinear optics (NLO) is the branch of optics that describes the behaviour of light in nonlinear media, that is, media in which the polarization density P responds non-linearly to the electric field E of the light. The non-linearity is typically observed only at very high light intensities (when the electric field of the light is >108 V/m and thus comparable to the atomic electric field of ~1011 V/m) such as those provided by lasers. Above the Schwinger limit, the vacuum itself is expected to become nonlinear. In nonlinear optics, the superposition principle no longer holds.

<span class="mw-page-title-main">Stimulated emission</span> Release of a photon triggered by another

Stimulated emission is the process by which an incoming photon of a specific frequency can interact with an excited atomic electron, causing it to drop to a lower energy level. The liberated energy transfers to the electromagnetic field, creating a new photon with a frequency, polarization, and direction of travel that are all identical to the photons of the incident wave. This is in contrast to spontaneous emission, which occurs at a characteristic rate for each of the atoms/oscillators in the upper energy state regardless of the external electromagnetic field.

In physics, attenuation is the gradual loss of flux intensity through a medium. For instance, dark glasses attenuate sunlight, lead attenuates X-rays, and water and air attenuate both light and sound at variable attenuation rates.

<span class="mw-page-title-main">Photoluminescence</span> 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.

<span class="mw-page-title-main">Raman scattering</span> Inelastic scattering of photons by matter

In physics, Raman scattering or the Raman effect is the inelastic scattering of photons by matter, meaning that there is both an exchange of energy and a change in the light's direction. Typically this effect involves vibrational energy being gained by a molecule as incident photons from a visible laser are shifted to lower energy. This is called normal Stokes-Raman scattering.

<span class="mw-page-title-main">Spontaneous parametric down-conversion</span> Optical process

Spontaneous parametric down-conversion is a nonlinear instant optical process that converts one photon of higher energy, into a pair of photons of lower energy, in accordance with the law of conservation of energy and law of conservation of momentum. It is an important process in quantum optics, for the generation of entangled photon pairs, and of single photons.

<span class="mw-page-title-main">Two-photon excitation microscopy</span> Fluorescence imaging technique

Two-photon excitation microscopy is a fluorescence imaging technique that is particularly well-suited to image scattering living tissue of up to about one millimeter in thickness. Unlike traditional fluorescence microscopy, where the excitation wavelength is shorter than the emission wavelength, two-photon excitation requires simultaneous excitation by two photons with longer wavelength than the emitted light. The laser is focused onto a specific location in the tissue and scanned across the sample to sequentially produce the image. Due to the non-linearity of two-photon excitation, mainly fluorophores in the micrometer-sized focus of the laser beam are excited, which results in the spatial resolution of the image. This contrasts with confocal microscopy, where the spatial resolution is produced by the interaction of excitation focus and the confined detection with a pinhole.

Quantum-cascade lasers (QCLs) are semiconductor lasers that emit in the mid- to far-infrared portion of the electromagnetic spectrum and were first demonstrated by Jérôme Faist, Federico Capasso, Deborah Sivco, Carlo Sirtori, Albert Hutchinson, and Alfred Cho at Bell Laboratories in 1994.

Ultrafast laser spectroscopy is a category of spectroscopic techniques using ultrashort pulse lasers for the study of dynamics on extremely short time scales. Different methods are used to examine the dynamics of charge carriers, atoms, and molecules. Many different procedures have been developed spanning different time scales and photon energy ranges; some common methods are listed below.

<span class="mw-page-title-main">Magneto-optical trap</span> Apparatus for trapping and cooling neutral atoms

In atomic, molecular, and optical physics, a magneto-optical trap (MOT) is an apparatus which uses laser cooling and a spatially-varying magnetic field to create a trap which can produce samples of cold, neutral atoms. Temperatures achieved in a MOT can be as low as several microkelvin, depending on the atomic species, which is two or three times below the photon recoil limit. However, for atoms with an unresolved hyperfine structure, such as 7Li, the temperature achieved in a MOT will be higher than the Doppler cooling limit.

<span class="mw-page-title-main">Silicon photonics</span> Photonic systems which use silicon as an optical medium

Silicon photonics is the study and application of photonic systems which use silicon as an optical medium. The silicon is usually patterned with sub-micrometre precision, into microphotonic components. These operate in the infrared, most commonly at the 1.55 micrometre wavelength used by most fiber optic telecommunication systems. The silicon typically lies on top of a layer of silica in what is known as silicon on insulator (SOI).

Photoelectrochemical processes are processes in photoelectrochemistry; they usually involve transforming light into other forms of energy. These processes apply to photochemistry, optically pumped lasers, sensitized solar cells, luminescence, and photochromism.

A parametric process is an optical process in which light interacts with matter in such a way as to leave the quantum state of the material unchanged. As a direct consequence of this there can be no net transfer of energy, momentum, or angular momentum between the optical field and the physical system. In contrast a non-parametric process is a process in which any part of the quantum state of the system changes.

Two-photon photovoltaic effect is an energy collection method based on two-photon absorption (TPA). The TPP effect can be thought of as the nonlinear equivalent of the traditional photovoltaic effect involving high optical intensities. This effect occurs when two photons are absorbed at the same time resulting in an electron-hole pair.

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.

<span class="mw-page-title-main">Semiconductor laser theory</span> 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.

<span class="mw-page-title-main">Two-photon circular dichroism</span>

Two-photon circular dichroism (TPCD), the nonlinear counterpart of electronic circular dichroism (ECD), is defined as the differences between the two-photon absorption (TPA) cross-sections obtained using left circular polarized light and right circular polarized light.

Stimulated Raman spectroscopy, also referred to as stimulated Raman scattering (SRS) is a form of spectroscopy employed in physics, chemistry, biology, and other fields. The basic mechanism resembles that of spontaneous Raman spectroscopy: a pump photon, of the angular frequency , which is scattered by a molecule has some small probability of inducing some vibrational transition, as opposed to inducing a simple Rayleigh transition. This makes the molecule emit a photon at a shifted frequency. However, SRS, as opposed to spontaneous Raman spectroscopy, is a third-order non-linear phenomenon involving a second photon—the Stokes photon of angular frequency —which stimulates a specific transition. When the difference in frequency between both photons resembles that of a specific vibrational transition the occurrence of this transition is resonantly enhanced. In SRS, the signal is equivalent to changes in the intensity of the pump and Stokes beams. The signals are typically rather low, of the order of a part in 10^5, thus calling for modulation-transfer techniques: one beam is modulated in amplitude and the signal is detected on the other beam via a lock-in amplifier. Employing a pump laser beam of a constant frequency and a Stokes laser beam of a scanned frequency allows for the unraveling of the spectral fingerprint of the molecule. This spectral fingerprint differs from those obtained by other spectroscopy methods such as Rayleigh scattering as the Raman transitions confer to different exclusion rules than those that apply for Rayleigh transitions.

<span class="mw-page-title-main">Photoacoustic microscopy</span>

Photoacoustic microscopy is an imaging method based on the photoacoustic effect and is a subset of photoacoustic tomography. Photoacoustic microscopy takes advantage of the local temperature rise that occurs as a result of light absorption in tissue. Using a nanosecond pulsed laser beam, tissues undergo thermoelastic expansion, resulting in the release of a wide-band acoustic wave that can be detected using a high-frequency ultrasound transducer. Since ultrasonic scattering in tissue is weaker than optical scattering, photoacoustic microscopy is capable of achieving high-resolution images at greater depths than conventional microscopy methods. Furthermore, photoacoustic microscopy is especially useful in the field of biomedical imaging due to its scalability. By adjusting the optical and acoustic foci, lateral resolution may be optimized for the desired imaging depth.

<span class="mw-page-title-main">Non-degenerate two-photon absorption</span> Simultaneous absorption of two photons of differing energies by a molecule

In atomic physics, non-degenerate two-photon absorption or two-color two-photon excitation is a type of two-photon absorption (TPA) where two photons with different energies are (almost) simultaneously absorbed by a molecule, promoting a molecular electronic transition from a lower energy state to a higher energy state. The sum of the energies of the two photons is equal to, or larger than, the total energy of the transition.

References

  1. Tkachenko, Nikolai V. (2006). "Appendix C. Two photon absorption". Optical Spectroscopy: Methods and Instrumentations. Elsevier. p. 293. ISBN   978-0-08-046172-4.
  2. Goeppert-Mayer M (1931). "Über Elementarakte mit zwei Quantensprüngen". Annals of Physics. 9 (3): 273–95. Bibcode:1931AnP...401..273G. doi: 10.1002/andp.19314010303 .
  3. Kaiser, W.; Garrett, C. G. B. (1961). "Two-Photon Excitation in CaF2:Eu2+". Physical Review Letters. 7 (6): 229. Bibcode:1961PhRvL...7..229K. doi:10.1103/PhysRevLett.7.229.
  4. Abella, I.D. (1962). "Optical double-quantum absorption in cesium vapor". Physical Review Letters. 9 (11): 453. Bibcode:1962PhRvL...9..453A. doi:10.1103/physrevlett.9.453.
  5. Braunstein, R.; Ockman, N. (20 April 1964). "Optical Double-Photon Absorption in CdS". Physical Review. 134 (2A): A499. Bibcode:1964PhRv..134..499B. doi:10.1103/PhysRev.134.A499.
  6. Hayat, Alex; Nevet, Amir; Ginzburg, Pavel; Orenstein, Meir (2011-08-01). "Applications of two-photon processes in semiconductor photonic devices: invited review". Semiconductor Science and Technology. 26 (8): 083001. Bibcode:2011SeScT..26h3001H. doi:10.1088/0268-1242/26/8/083001. ISSN   0268-1242. S2CID   51993416.
  7. Xu, Chris; Webb, Watt (1996). "Measurement of two-photon excitation cross sections of molecular fluorophores with data from 690 to 1050 nm". JOSA B. 13 (3): 481–491. Bibcode:1996JOSAB..13..481X. doi:10.1364/JOSAB.13.000481.
  8. Trejo-Valdez, M.; Torres-Martínez, R.; Peréa-López, N.; Santiago-Jacinto, P.; Torres-Torres, C. (2010-06-10). "Contribution of the Two-Photon Absorption to the Third Order Nonlinearity of Au Nanoparticles Embedded in TiO2 Films and in Ethanol Suspension". The Journal of Physical Chemistry C. 114 (22): 10108–10113. doi:10.1021/jp101050p. ISSN   1932-7447.
  9. Mahr, H. (2012). "Chapter 4. Two-Photon Absorption Spectroscopy". In Herbert Rabin, C. L. Tang (ed.). Quantum Electronics: A Treatise, Volume 1. Nonlinear Optics, Part A. Academic Press. pp. 286–363. ISBN   978-0-323-14818-4.
  10. Xu, Chris; Webb, Watt (1996). "Measurement of two-photon excitation cross sections of molecular fluorophores with data from 690 to 1050 nm". JOSA B. 13 (3): 481–491. Bibcode:1996JOSAB..13..481X. doi:10.1364/JOSAB.13.000481.
  11. Powerpoint presentation http://www.chem.ucsb.edu/~ocf/lecture_ford.ppt
  12. Hayat, Alex; Nevet, Amir; Ginzburg, Pavel; Orenstein, Meir (2011). "Applications of two-photon processes in semiconductor photonic devices: Invited review". Semiconductor Science and Technology. 26 (8): 083001. Bibcode:2011SeScT..26h3001H. doi:10.1088/0268-1242/26/8/083001. S2CID   51993416.
  13. Dorfer, Christian; Hits, Dmitry; Kasmi, Lamia; Kramberger, Gregor (2019). "Three-dimensional charge transport mapping by two-photon absorption edge transient-current technique in synthetic single-crystalline diamond". Applied Physics Letters. 114 (20): 203504. arXiv: 1905.09648 . Bibcode:2019ApPhL.114t3504D. doi:10.1063/1.5090850. hdl: 11311/1120457 . S2CID   165163659.
  14. Wu, En-Shinn; Strickler, James H.; Harrell, W. R.; Webb, Watt W. (1992-06-01). Cuthbert, John D. (ed.). "Two-photon lithography for microelectronic application". Optical/Laser Microlithography V. SPIE. 1674: 776–782. Bibcode:1992SPIE.1674..776W. doi:10.1117/12.130367. S2CID   135759224.
  15. Maruo, Shoji; Nakamura, Osamu; Kawata, Satoshi (1997-01-15). "Three-dimensional microfabrication with two-photon-absorbed photopolymerization". Optics Letters. 22 (2): 132–134. Bibcode:1997OptL...22..132M. doi:10.1364/OL.22.000132. ISSN   1539-4794. PMID   18183126.
  16. Gissibl, Timo; Thiele, Simon; Herkommer, Alois; Giessen, Harald (August 2016). "Two-photon direct laser writing of ultracompact multi-lens objectives". Nature Photonics. 10 (8): 554–560. Bibcode:2016NaPho..10..554G. doi:10.1038/nphoton.2016.121. ISSN   1749-4893. S2CID   49191430.
  17. Jaiswal, Arun; Rastogi, Chandresh Kumar; Rani, Sweta; Singh, Gaurav Pratap; Saxena, Sumit; Shukla, Shobha (2023-04-21). "Two decades of two-photon lithography: Materials science perspective for additive manufacturing of 2D/3D nano-microstructures". iScience. 26 (4): 106374. Bibcode:2023iSci...26j6374J. doi:10.1016/j.isci.2023.106374. ISSN   2589-0042. PMC   10121806 . PMID   37096047.
  18. Galanopoulos, Stratos; Chatzidai, Nikoleta; Melissinaki, Vasileia; Selimis, Alexandros; Schizas, Charalampos; Farsari, Maria; Karalekas, Dimitris (September 2014). "Design, Fabrication and Computational Characterization of a 3D Micro-Valve Built by Multi-Photon Polymerization". Micromachines. 5 (3): 505–514. doi: 10.3390/mi5030505 . ISSN   2072-666X.
  19. Yang, Liang; Mayer, Frederik; Bunz, Uwe H. F.; Blasco, Eva; Wegener, Martin (2021). "Multi-material multi-photon 3D laser micro- and nanoprinting". Light: Advanced Manufacturing. 2: 1. doi:10.37188/lam.2021.017 (inactive 2024-03-04). ISSN   2689-9620.{{cite journal}}: CS1 maint: DOI inactive as of March 2024 (link)
  20. Raimondi, Manuela T.; Eaton, Shane M.; Nava, Michele M.; Laganà, Matteo; Cerullo, Giulio; Osellame, Roberto (2012-06-26). "Two-photon laser polymerization: from fundamentals to biomedical application in tissue engineering and regenerative medicine". Journal of Applied Biomaterials & Functional Materials. 10 (1): 55–65. doi:10.5301/JABFM.2012.9278. ISSN   2280-8000. PMID   22562455.
  21. Izquierdo-Serra, Mercè; Gascón-Moya, Marta; Hirtz, Jan J.; Pittolo, Silvia; Poskanzer, Kira E.; Ferrer, Èric; Alibés, Ramon; Busqué, Félix; Yuste, Rafael; Hernando, Jordi; Gorostiza, Pau (2014-06-18). "Two-Photon Neuronal and Astrocytic Stimulation with Azobenzene-Based Photoswitches". Journal of the American Chemical Society. 136 (24): 8693–8701. doi:10.1021/ja5026326. ISSN   0002-7863. PMC   4096865 . PMID   24857186.
  22. Carroll, Elizabeth C.; Berlin, Shai; Levitz, Joshua; Kienzler, Michael A.; Yuan, Zhe; Madsen, Dorte; Larsen, Delmar S.; Isacoff, Ehud Y. (2015-02-17). "Two-photon brightness of azobenzene photoswitches designed for glutamate receptor optogenetics". Proceedings of the National Academy of Sciences. 112 (7): E776-85. Bibcode:2015PNAS..112E.776C. doi: 10.1073/pnas.1416942112 . ISSN   0027-8424. PMC   4343171 . PMID   25653339.
  23. Pittolo, Silvia; Lee, Hyojung; Lladó, Anna; Tosi, Sébastien; Bosch, Miquel; Bardia, Lídia; Gómez-Santacana, Xavier; Llebaria, Amadeu; Soriano, Eduardo; Colombelli, Julien; Poskanzer, Kira E.; Perea, Gertrudis; Gorostiza, Pau (2019-07-02). "Reversible silencing of endogenous receptors in intact brain tissue using 2-photon pharmacology". Proceedings of the National Academy of Sciences. 116 (27): 13680–13689. Bibcode:2019PNAS..11613680P. doi: 10.1073/pnas.1900430116 . ISSN   0027-8424. PMC   6613107 . PMID   31196955.
  24. Riefolo, Fabio; Matera, Carlo; Garrido-Charles, Aida; Gomila, Alexandre M. J.; Sortino, Rosalba; Agnetta, Luca; Claro, Enrique; Masgrau, Roser; Holzgrabe, Ulrike; Batlle, Montserrat; Decker, Michael; Guasch, Eduard; Gorostiza, Pau (2019-05-08). "Optical Control of Cardiac Function with a Photoswitchable Muscarinic Agonist". Journal of the American Chemical Society. 141 (18): 7628–7636. doi:10.1021/jacs.9b03505. hdl: 2445/147236 . ISSN   0002-7863. PMID   31010281. S2CID   128361100.
  25. Cabré, Gisela; Garrido-Charles, Aida; Moreno, Miquel; Bosch, Miquel; Porta-de-la-Riva, Montserrat; Krieg, Michael; Gascón-Moya, Marta; Camarero, Núria; Gelabert, Ricard; Lluch, José M.; Busqué, Félix; Hernando, Jordi; Gorostiza, Pau; Alibés, Ramon (2019-02-22). "Rationally designed azobenzene photoswitches for efficient two-photon neuronal excitation". Nature Communications. 10 (1): 907. Bibcode:2019NatCo..10..907C. doi:10.1038/s41467-019-08796-9. ISSN   2041-1723. PMC   6385291 . PMID   30796228.
  26. Kellner, Shai; Berlin, Shai (January 2020). "Two-Photon Excitation of Azobenzene Photoswitches for Synthetic Optogenetics". Applied Sciences. 10 (3): 805. doi: 10.3390/app10030805 . ISSN   2076-3417.
  27. Kogej, T.; Beljonne, D.; Meyers, F.; Perry, J.W.; Marder, S.R.; Brédas, J.L. (1998). "Mechanisms for enhancement of two-photon absorption in donor–acceptor conjugated chromophores". Chemical Physics Letters. 298 (1): 1–6. Bibcode:1998CPL...298....1K. doi:10.1016/S0009-2614(98)01196-8.
  28. "Two Photon Absorption Spectra | KBFI KBFI". KBFI. Retrieved 2020-09-03.
  29. "Two-photon action cross-sections".
  30. Kamada, Kenji; Ohta, Koji; Kubo, Takashi; Shimizu, Akihiro; Morita, Yasushi; Nakasuji, Kazuhiro; Kishi, Ryohei; Ohta, Suguru; Furukawa, Shin-Ichi; Takahashi, Hideaki; Nakano, Masayoshi (2007). "Strong Two-Photon Absorption of Singlet Diradical Hydrocarbons". Angewandte Chemie International Edition. 46 (19): 3544–3546. doi:10.1002/anie.200605061. PMID   17385813.
  31. Bass, Michael (1994). HANDBOOK OF OPTICS Volume I. McGraw-Hill Professional; 2 edition (September 1, 1994). 9 .32. ISBN   978-0-07-047740-7.
  32. Marvin, Weber (2003). Handbook of optical materials. Laser and Optical Science and Technology Series. The CRC Press. APPENDIX V. ISBN   978-0-8493-3512-9.
  33. 1 2 3 Dragonmir, Adrian; McInerney, John G.; Nikogosyan, David N. (2002). "Femtosecond Measurements of Two-Photon Absorption Coefficients at λ = 264 nm in Glasses, Crystals, and Liquids". Applied Optics. 41 (21): 4365–4376. Bibcode:2002ApOpt..41.4365D. doi:10.1364/AO.41.004365. PMID   12148767.
  34. Nikogosyan, D. N.; Angelov, D. A. (1981). "Formation of free radicals in water under high-power laser UV irradiation". Chemical Physics Letters. 77 (1): 208–210. Bibcode:1981CPL....77..208N. doi:10.1016/0009-2614(81)85629-1.
  35. Underwood, J.; Wittig, C. (2004). "Two photon photodissociation of H2O via the B state". Chemical Physics Letters. 386 (1): 190–195. Bibcode:2004CPL...386..190U. doi:10.1016/j.cplett.2004.01.030.
  36. Chluba, J.; Sunyaev, R. A. (2006). "Induced two-photon decay of the 2s level and the rate of cosmological hydrogen recombination". Astronomy and Astrophysics. 446 (1): 39–42. arXiv: astro-ph/0508144 . Bibcode:2006A&A...446...39C. doi:10.1051/0004-6361:20053988. S2CID   119526307.
  37. Spitzer, L.; Greenstein, J. (1951). "Continuous emission from planetary nebulae". Astrophysical Journal. 114: 407. Bibcode:1951ApJ...114..407S. doi:10.1086/145480.
  38. Gurzadyan, G. A. (1976). "Two-photon emission in planetary nebula IC 2149". Publications of the Astronomical Society of the Pacific. 88 (526): 891–895. doi:10.1086/130041. JSTOR   40676041.
  39. Hayat, A.; Ginzburg, P.; Orenstein, M. (2008). "Observation of Two-Photon Emission from Semiconductors". Nature Photonics. 2 (4): 238. doi:10.1038/nphoton.2008.28.