Surface second harmonic generation

Last updated

Surface second harmonic generation is a method for probing interfaces in atomic and molecular systems. In second harmonic generation (SHG), the light frequency is doubled, essentially converting two photons of the original beam of energy E into a single photon of energy 2E as it interacts with noncentrosymmetric media. Surface second harmonic generation is a special case of SHG where the second beam is generated because of a break of symmetry caused by an interface. Since centrosymmetric symmetry in centrosymmetric media is only disrupted in the first (occasionally second and third) atomic or molecular layer of a system, properties of the second harmonic signal then provide information about the surface atomic or molecular layers only. Surface SHG is possible even for materials which do not exhibit SHG in the bulk. [1] Although in many situations the dominant second harmonic signal arises from the broken symmetry at the surface, the signal in fact always has contributions from both the surface and bulk. [2] Thus, the most sensitive experiments typically involve modification of a surface and study of the subsequent modification of the harmonic generation properties.

Contents

History

Second harmonic generation from a surface was first observed by Terhune, Maker, and Savage at the Ford Motor Company in 1962, [3] one year after Franken et al. first discovered second harmonic generation in bulk crystals. Prior to Terhune's discovery, it was believed that crystals could only exhibit second harmonic generation if the crystal was noncentrosymmetric. Terhune observed that calcite, a centrosymmetric crystal which is only capable of SHG in the bulk in the presence of an applied electric field which would break the symmetry of the electronic structure, surprisingly also produced a second harmonic signal in the absence of an external electric field. During the 1960s, SHG was observed for many other centrosymmetric media including metals, semiconductors, oxides, and liquids. In 1968, Bloembergen et al. [1] showed that the second harmonic signal was generated from the surface. Interest in this field waned during the 1970s and only a handful of research groups investigated surface SHG, most notably Y. R. Shen's group at University of California at Berkeley. [4] [5] During the 70s and 80s, most of the research in this field focused on understanding the electronic response, particularly in metals. In 1981, Chen et al. showed that SHG could be used to detect individual monolayers, [6] and since then, much research has gone into using and understanding SHG as surface probe of molecular adsorption and orientation. [7]

Excitation of second harmonic signal

Just as bulk second harmonic generation, surface SHG arises out of the second-order susceptibility tensor χ(2). While the χ(2) tensor contains 27 elements, many of these elements are reduced by symmetry arguments. The exact nature of these arguments depends on the application. When determining molecular orientation, it is assumed that χ(2) is rotationally invariant around the z-axis (normal to the surface). The number of tensor elements reduces from 27 to the following 7 independent quantities: χZZZ, χZXX = χZYY, χXZX = χYZY, χXXZ = χYYZ, χXYZ = -χYXZ, χXZY = -χYZX, χZXY = -χZYX. Second Harmonic Generation further restricts the independent terms by requiring the tensor is symmetric in the last two indices reducing the number of independent tensor terms to 4: χZZZ, χZXX (equivalently χZYY), χXXZ (equivalently χXZX, χYZY, χYYZ), χXYZ (equivalently χXZY, -χYXZ, -χYZX). In order for χZXY = -χZYX to hold under this final condition, both terms must be 0. The four independent terms are material dependent properties and can vary as the external conditions change. These four terms give rise to the second harmonic signal, and allow for calculation of material properties such as electronic structure, atomic organization, and molecular orientation. Detailed analysis of the second harmonic generation from surfaces and interfaces, as well as the ability to detect monolayers and sub-monolayers, may be found in Guyot-Sionnest et al. [8]

Applications

Interface structure

Figure 1: Crystal Surface SHG Setup SHG Metal Anisotropy.GIF
Figure 1: Crystal Surface SHG Setup
Figure 2: Polar Crystal Surface SHG Response (arbitrary units)(adapted from ) Fcc p polarized.gif
Figure 2: Polar Crystal Surface SHG Response (arbitrary units)(adapted from )

It may seem paradoxical at first that surface SHG which relies on a break in symmetry is possible in crystals which have an inherent symmetric structure. At a crystalline interface half of the atomic forces experienced in the bulk crystal are not present which causes changes in the atomic and electronic structures. There are two major changes that occur at the interface: 1) the interplanar distances of the top layers change and 2) the atoms redistribute themselves to a completely new packing structure. While symmetry is maintained in the surface planes, the break in symmetry out-of-plane modifies the second-order susceptibility tensor χ(2), giving rise to optical second harmonic generation. Typical measurements of SHG from crystalline surfaces structures are performed by rotating the sample in an incident beam (Figure 1). The second harmonic signal will vary with the azimuth angle of the sample due to the symmetry of the atomic and electronic structure (Figure 2). As a result, surface SHG theory is highly dependent on geometry of the superstructure. Since electron interactions are responsible for the SHG response, the jellium model is usually numerically solved using Density Functional Theory to predict the SHG response of a given surface. [10] SHG sensitivity to surface structure approach was effectively demonstrated by Heinz, Loy, and Thompson, working for IBM in 1985. [11] They showed that the SHG signal from a freshly cleaved Si(111) surface would alter its behavior as the temperature was raised and the superstructure changed from a 2x1 structure to the 7x7 structure. Noting the change in signal, they were able to verify the existence of one mirror plane in the 2x1 construction and 3 mirror planes in the 7x7 construction thereby providing new information to the bonding structure of the surface atoms. Since then, surface SHG has been used to probe many other metallic surfaces such as reconstructed gold (110), [12] Pd(111), [13] and Al(100). [14]

Perhaps one of the most powerful uses of surface SHG is the probing of surface structure of buried interfaces. Traditional surface tools such as atomic force microscopy and scanning tunneling microscopy as well as many forms of electron diffraction must be conducted under vacuum, and are not sensitive to interfaces deeper in the probed medium. SHG measurements allow the incident laser beam to pass without interaction through higher level materials to the target interface where the second harmonic signal is generated. In cases where the transmitting materials do interact with the beam, these contributions to the second harmonic signal can be resolved in other experiments and subtracted out. The resulting measured second harmonic signal contains the second harmonic component from the buried interface alone. This type of measurement is useful for determining the surface structure of the interface. As an example, Cheikh-Rouhou et al. demonstrated this process to resolve interface structures of 5 layer systems. [15]

Adsorption measurements

Figure 3: Surface SHG Adsorption Isotherm for Rhodamine 6G (adapted from SHG R6G.gif
Figure 3: Surface SHG Adsorption Isotherm for Rhodamine 6G (adapted from

Surface SHG is useful for monitoring the growth of monolayers on a surface. As particles adsorb, the SHG signal is altered. Two common applications in surface science are the adsorption of small gas molecules onto a surface and the adsorption of dissolved dye molecules in a liquid to a surface. Bourguignon et al. [13] showed that as carbon monoxide is adsorbed onto a Pd(111) surface, the SHG signal decreased exponentially as predicted by the Langmuir isotherm. As CO coverage approached 1 monolayer, the SHG intensity leveled off. Larger molecules like dyes often can form multilayers on a surface, and this can be measured in situ using SHG. As the first monolayer forms, the intensity can often be seen to increase to a maximum until a uniform distribution of particles is obtained (Figure 3). As additional particles adsorb and the second monolayer begins to form, the SHG signal decreases until it reaches a minimum at the completion of the second monolayer. This alternating behavior can typically be seen for the growth of monolayers. [4] [16] As additional layers form, the SHG response of the substrate is screened by the adsorbate and eventually, the SHG signal levels off.

Molecular orientation

Figure 4: Total internal reflection geometry of surface SHG Surface SHG.GIF
Figure 4: Total internal reflection geometry of surface SHG

As molecular layers adsorb to surfaces it is often useful to know the molecular orientation of the adsorbed molecules. Molecular orientation can be probed by observing the polarization of the second harmonic signal, generated from a polarized beam. Figure 4 shows a typical experimental geometry for molecular orientation experiments. The beam is incident on the sample in a total internal reflection geometry which improves the second harmonic signal because as the wave propagates along the interface, additional second harmonic photons are generated, [1] By rotating either the polarizer or the analyzer, the s- and p-polarized signals are measured which allow for the calculation of the second-order susceptibility tensor χ(2). Simpson's research group has studied this phenomenon in depth. [17] [18] [19] The molecular orientation can differ from the laboratory axis in three directions, corresponding to three angles. Typically, SHG measurements of this type are only able to extract a single parameter, namely the molecular orientation with respect to the surface normal.

Calculation of molecular orientation

When dealing with adsorbed molecules on a surface, it is typical to find a uniaxial distribution of the molecules, resulting in x- and y- coordinate terms to be interchangeable. When analyzing the second-order susceptibility tensor χ(2), the quantities χXYZ = -χYXZ must be 0 and only three independent tensor terms remain: χzzz, χzxx, and χxxz. The intensities of the s and p polarizations in the second harmonic are given by following relationships: [18]

where γ is the polarization angle with γ = 0 corresponding to p-polarized light. The si terms depend on the experimental geometry are functions of the total internal reflection angles of the incident and second harmonic beams and the linear and nonlinear Fresnel factors respectively which relate the electric field components at the interface to incident and detected fields.

The second-order susceptibility tensor, χ(2), is the parameter which can be measured in second order experiments, but it does not explicitly provide insight to the molecular orientation of surface molecules. To determine molecular orientation, the second-order hyperpolarizability tensor β, must be calculated. For adsorbed molecules in a uniaxial distribution, the only independent hyperpolarizability tensor terms are βz’z’z’, βz’x’x’, and βx’x’z’ where ’ terms denote the molecular coordinate system as opposed to the laboratory coordinate system. β can be related to χ(2) through orientational averages. As an example, in an isotropic distribution on the surface, χ(2) elements are given by. [7]

where Ns is the surface number density of the adsorbed molecules, θ and Ψ are orientational angles relating the molecular coordinate system to the laboratory coordinate system, and <x> represents the average value of x. In many cases, only one or two of the molecular hyperpolarizability tensor are dominant. In these cases, the relationships between χ and β can be simplified. Bernhard Dick presents several of these simplifications. [20]

Additional applications

In addition to these applications, Surface SHG is used to probe other effects. [5] In surface spectroscopy, where either the fundamental or second harmonic are resonant with electronic transitions in the surface atoms, details can be determined about the electronic structure and band gaps. In monolayer microscopy the second harmonic signal is magnified and surface features are imaged with a resolution on the order of a wavelength. Surface SHG can also be used to monitor chemical reactions at a surface with picosecond resolution.

Related Research Articles

Frequency modulation synthesis Form of sound synthesis

Frequency modulation synthesis is a form of sound synthesis whereby the frequency of a waveform is changed by modulating its frequency with a modulator. The frequency of an oscillator is altered "in accordance with the amplitude of a modulating signal".

In electrodynamics, elliptical polarization is the polarization of electromagnetic radiation such that the tip of the electric field vector describes an ellipse in any fixed plane intersecting, and normal to, the direction of propagation. An elliptically polarized wave may be resolved into two linearly polarized waves in phase quadrature, with their polarization planes at right angles to each other. Since the electric field can rotate clockwise or counterclockwise as it propagates, elliptically polarized waves exhibit chirality.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition. By definition, a rotation about the origin is a transformation that preserves the origin, Euclidean distance, and orientation. Every non-trivial rotation is determined by its axis of rotation and its angle of rotation. Composing two rotations results in another rotation; every rotation has a unique inverse rotation; and the identity map satisfies the definition of a rotation. Owing to the above properties, the set of all rotations is a group under composition. Rotations are not commutative, making it a nonabelian group. Moreover, the rotation group has a natural structure as a manifold for which the group operations are smoothly differentiable; so it is in fact a Lie group. It is compact and has dimension 3.

Spherical harmonics

In mathematics and physical science, spherical harmonics are special functions defined on the surface of a sphere. They are often employed in solving partial differential equations in many scientific fields.

Poincaré half-plane model Upper-half plane model of hyperbolic non-Euclidean geometry

In non-Euclidean geometry, the Poincaré half-plane model is the upper half-plane, denoted below as H, together with a metric, the Poincaré metric, that makes it a model of two-dimensional hyperbolic geometry.

Bloch sphere Geometrical representation of the pure state space of a two-level quantum mechanical system

In quantum mechanics and computing, the Bloch sphere is a geometrical representation of the pure state space of a two-level quantum mechanical system (qubit), named after the physicist Felix Bloch.

The classical XY model is a lattice model of statistical mechanics. In general, the XY model can be seen as a specialization of Stanley's n-vector model for n = 2.

The rigid rotor is a mechanical model of rotating systems. An arbitrary rigid rotor is a 3-dimensional rigid object, such as a top. To orient such an object in space requires three angles, known as Euler angles. A special rigid rotor is the linear rotor requiring only two angles to describe, for example of a diatomic molecule. More general molecules are 3-dimensional, such as water, ammonia, or methane. The rigid-rotor Schroedinger equation is discussed in Section 11.2 on pages 240-253 of the textbook by Bunker and Jensen.

Transverse isotropy

A transversely isotropic material is one with physical properties that are symmetric about an axis that is normal to a plane of isotropy. This transverse plane has infinite planes of symmetry and thus, within this plane, the material properties are the same in all directions. Hence, such materials are also known as "polar anisotropic" materials. In geophysics, vertically transverse isotropy (VTI) is also known as radial anisotropy.

In quantum chemistry, the potential energy surfaces are obtained within the adiabatic or Born–Oppenheimer approximation. This corresponds to a representation of the molecular wave function where the variables corresponding to the molecular geometry and the electronic degrees of freedom are separated. The non separable terms are due to the nuclear kinetic energy terms in the molecular Hamiltonian and are said to couple the potential energy surfaces. In the neighbourhood of an avoided crossing or conical intersection, these terms cannot be neglected. One therefore usually performs one unitary transformation from the adiabatic representation to the so-called diabatic representation in which the nuclear kinetic energy operator is diagonal. In this representation, the coupling is due to the electronic energy and is a scalar quantity which is significantly easier to estimate numerically.

The magic angle is a precisely defined angle, the value of which is approximately 54.7356°. The magic angle is a root of a second-order Legendre polynomial, P2(cos θ) = 0, and so any interaction which depends on this second-order Legendre polynomial vanishes at the magic angle. This property makes the magic angle of particular importance in magic angle spinning solid-state NMR spectroscopy. In magnetic resonance imaging, structures with ordered collagen, such as tendons and ligaments, oriented at the magic angle may appear hyperintense in some sequences; this is called the magic angle artifact or effect.

Second-harmonic generation

Second-harmonic generation is a nonlinear optical process in which two photons with the same frequency interact with a nonlinear material, are "combined", and generate a new photon with twice the energy of the initial photons, that conserves the coherence of the excitation. It is a special case of sum-frequency generation, and more generally of harmonic generation.

The Wigner D-matrix is a unitary matrix in an irreducible representation of the groups SU(2) and SO(3). The complex conjugate of the D-matrix is an eigenfunction of the Hamiltonian of spherical and symmetric rigid rotors. The matrix was introduced in 1927 by Eugene Wigner. D stands for Darstellung, which means "representation" in German.

Gravitational lensing formalism

In general relativity, a point mass deflects a light ray with impact parameter by an angle approximately equal to

This article derives the main properties of rotations in 3-dimensional space.

Amplitude amplification is a technique in quantum computing which generalizes the idea behind the Grover's search algorithm, and gives rise to a family of quantum algorithms. It was discovered by Gilles Brassard and Peter Høyer in 1997, and independently rediscovered by Lov Grover in 1998.

Sum frequency generation spectroscopy (SFG) is a nonlinear laser spectroscopy technique used to analyze surfaces and interfaces. In a typical SFG setup, two laser beams mix at an interface and generate an output beam with a frequency equal to the sum of the two input frequencies, traveling in a direction given by the sum of the incident beams' wavevectors. The technique was developed in 1987 by Yuen-Ron Shen and his students as an extension of second harmonic generation spectroscopy and rapidly applied to deduce the composition, orientation distributions, and structural information of molecules at gas–solid, gas–liquid and liquid–solid interfaces. Soon after its invention, Philippe Guyot-Sionnest extended the technique to obtain the first measurements of electronic and vibrational dynamics at surfaces. SFG has advantages in its ability to be monolayer surface sensitive, ability to be performed in situ, and its capability to provide ultrafast time resolution. SFG gives information complementary to infrared and Raman spectroscopy.

Optical phase space phase space used in quantum optics

In quantum optics, an optical phase space is a phase space in which all quantum states of an optical system are described. Each point in the optical phase space corresponds to a unique state of an optical system. For any such system, a plot of the quadratures against each other, possibly as functions of time, is called a phase diagram. If the quadratures are functions of time then the optical phase diagram can show the evolution of a quantum optical system with time.

For many paramagnetic materials, the magnetization of the material is directly proportional to an applied magnetic field, for large temperatures small fields. However, if the material is heated, this proportionality is reduced. For a fixed value of the field, the magnetic susceptibility is inversely proportional to temperature, that is

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

References

  1. 1 2 3 Bloembergen, N.; Chang, R. K.; Jha, S. S.; Lee, C. H. (1968-10-15). "Optical Second-Harmonic Generation in Reflection from Media with Inversion Symmetry". Physical Review. American Physical Society (APS). 174 (3): 813–822. doi:10.1103/physrev.174.813. ISSN   0031-899X.
  2. Guyot-Sionnest, P.; Shen, Y.R.;"Bulk contribution in surface second-harmonic generation". Physical Review B, 38, 12, 1988 p 7985-7989. https://doi.org/10.1103/PhysRevB.38.7985
  3. Terhune, R. W.; Maker, P. D.; Savage, C. M. (1962). "Optical Harmonic Generation in Calcite". Physical Review Letters. 8 (10): 404–406. doi:10.1103/PhysRevLett.8.404. ISSN   0031-9007.
  4. 1 2 Shen, Y. R. (1986). "Surface Second Harmonic Generation: A New Technique for Surface Studies". Annual Review of Materials Science . Annual Reviews. 16 (1): 69–86. doi:10.1146/annurev.ms.16.080186.000441. ISSN   0084-6600.
  5. 1 2 Shen, Y R (1989). "Optical Second Harmonic Generation at Interfaces". Annual Review of Physical Chemistry . Annual Reviews. 40 (1): 327–350. doi:10.1146/annurev.pc.40.100189.001551. ISSN   0066-426X.
  6. Chen, C. K.; Heinz, T. F.; Ricard, D.; Shen, Y. R. (1981-04-13). "Detection of Molecular Monolayers by Optical Second-Harmonic Generation". Physical Review Letters. American Physical Society (APS). 46 (15): 1010–1012. doi:10.1103/physrevlett.46.1010. ISSN   0031-9007.
  7. 1 2 Heinz, T. F. Nonlinear Surface Electromagnetic Phenomena; North-Holland: New York, 1991; Chapter 5
  8. Guyot-Sionnest, P.; Chen, C. K., Shen, Y. R. General considerations on optical second-harmonic generation from surfaces and interfaces Physical Review B , 33, 12, 1986 p 8254-8263. https://doi.org/10.1103/PhysRevB.33.8254
  9. Lohner, F.P.; Villaeys, A.A. (1998). "Anisotropy analysis of the SHG intensity by surfaces of simple metals". Optics Communications. Elsevier BV. 154 (4): 217–224. doi:10.1016/s0030-4018(98)00314-9. ISSN   0030-4018.
  10. Weber, M.; Liebsch, A. (1987-05-15). "Density-functional approach to second-harmonic generation at metal surfaces". Physical Review B. American Physical Society (APS). 35 (14): 7411–7416. doi:10.1103/physrevb.35.7411. ISSN   0163-1829. PMID   9941043.
  11. Heinz, T. F.; Loy, M. M. T.; Thompson, W. A. (1985-01-07). "Study of Si(111) Surfaces by Optical Second-Harmonic Generation: Reconstruction and Surface Phase Transformation". Physical Review Letters. American Physical Society (APS). 54 (1): 63–66. doi:10.1103/physrevlett.54.63. ISSN   0031-9007. PMID   10030885.
  12. Iwai, Tetsuya; Mizutani, Goro Shinku/Journal of the Vacuum Society of Japan47:171-174 (2004)
  13. 1 2 Bourguignon, Bernard; Zheng, Wanquan; Carrez, Serge; Fournier, Frédéric; Gaillard, Michel L.; Dubost, Henri (2002). "On the anisotropy and CO coverage dependence of SHG from Pd(111)". Surface Science. Elsevier BV. 515 (2–3): 567–574. doi:10.1016/s0039-6028(02)02000-9. ISSN   0039-6028.
  14. Jakobsen, C.; Podenas, D.; Pedersen, K. (1994). "Optical second-harmonic generation from vicinal Al(100) crystals". Surface Science. Elsevier BV. 321 (1–2): 1–7. doi:10.1016/0039-6028(94)90021-3. ISSN   0039-6028.
  15. Cheikh-Rouhou, W.; Sampaio, L.C.; Bartenlian, B.; Beauvillain, P.; Brun, A.; et al. (2002). "SHG anisotropy in Au/Co/Au/Cu/vicinal Si(111)". Journal of Magnetism and Magnetic Materials. Elsevier BV. 240 (1–3): 532–535. doi:10.1016/s0304-8853(01)00840-x. ISSN   0304-8853.
  16. 1 2 Kikteva, Tanya; Star, Dmitry; Leach, Gary W. (2000). "Optical Second Harmonic Generation Study of Malachite Green Orientation and Order at the Fused-Silica/Air Interface". The Journal of Physical Chemistry B. American Chemical Society (ACS). 104 (13): 2860–2867. doi:10.1021/jp992728b. ISSN   1520-6106.
  17. Simpson, Garth J.; Westerbuhr, Sarah G.; Rowlen, Kathy L. (2000). "Molecular Orientation and Angular Distribution Probed by Angle-Resolved Absorbance and Second Harmonic Generation". Analytical Chemistry. American Chemical Society (ACS). 72 (5): 887–898. doi:10.1021/ac9912956. ISSN   0003-2700. PMID   10739189.
  18. 1 2 Simpson, Garth J.; Rowlen, Kathy L. (2000). "Orientation-Insensitive Methodology for Second Harmonic Generation. 1. Theory". Analytical Chemistry. American Chemical Society (ACS). 72 (15): 3399–3406. doi:10.1021/ac000346s. ISSN   0003-2700. PMID   10952518.
  19. Simpson, Garth J.; Rowlen, Kathy L. (2000). "Orientation-Insensitive Methodology for Second Harmonic Generation. 2. Application to Adsorption Isotherm and Kinetics Measurements". Analytical Chemistry. American Chemical Society (ACS). 72 (15): 3407–3411. doi:10.1021/ac000347k. ISSN   0003-2700. PMID   10952519.
  20. Dick, Bernhard (1985). "Irreducible tensor analysis of sum- and difference-frequency generation in partially oriented samples" (PDF). Chemical Physics. Elsevier BV. 96 (2): 199–215. doi:10.1016/0301-0104(85)85085-0. ISSN   0301-0104.