Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Facet-dependent electric-field-induced second harmonic generation in silicon and zincblende

Open Access Open Access

Abstract

We discuss electric-field-induced second harmonic (EFISH) generation for silicon and zincblende facets (001), (011), and (111), employing the full fourth-rank tensor representation of the third-order susceptibility. Then we directly relate these 81 tensor elements with the contracted or Voigt matrix representation. Using group theory, we show that the number of independent elements is only two; however, at different facets, different linear combinations thereof appear. Also, specific expressions for the resulting s- and p-polarized second harmonic polarization are given for incident s- and p-polarizations, for the first time explaining the facet and angle of incidence dependence of EFISH. Finally, a classical oscillator model is used to explain the response of the electrons and the material combined with a direct physical interpretation of the breaking of the symmetry and thus the deformation of the electronic charge density along the bonds. Through this model we propose a connection between the strength parameter b for third harmonic generation and the second harmonic signal originated by EFISH mechanism.

© 2017 Optical Society of America

1. INTRODUCTION

Nonlinear phenomena have been widely studied since the invention of the laser by Townes et al. [1]. Subsequently, second harmonic generation (SHG) [2], third harmonic generation (THG) [35], higher harmonic generation (HHG) [69] and mixing frequency techniques [1013] have been intensively studied theoretically and experimentally.

Another nonlinear optical phenomenon, studied soon afterwards, is electric-field-induced second harmonic (EFISH) generation, first observed experimentally by Terhune and coworkers using calcite crystals [5]. Later, the dependence of SHG with respect to a DC electric field for silver and silicon surfaces was reported by Lee and coworkers [14], where they already observed that for a p-polarized light the EFISH effect is large; however, for s-polarized light, it is almost negligible.

Experimentally it is worth mentioning that even in inversion symmetric materials, where dipolar SHG is forbidden (χijkdipole=0), bulk SHG can still occur due to symmetry breaking of the applied DC field. Therefore, the argument that only surfaces—and not the bulk—of Si contribute to dipolar SHG in inversion symmetric materials is then only correct if no static electric field is present in the sample. EFISH describes the interaction of matter with two fundamental electromagnetic waves with the addition of a DC electric field; thus, these three electric fields generate together a new signal with the doubled frequency of the impinging beams. SHG, generated through EFISH, is therefore a third-order nonlinear phenomenon. Third-order coefficients for EFISH have been measured experimentally for gases, liquids [15] and glass [16]. Also, EFISH was deeply and intensively studied experimentally, as well as theoretically, by Aktsipetrov and coworkers for metal-oxide-semiconductors structures and Si-SiO2 interfaces [17], where in some cases it has been possible to separate the second harmonic (SH) signal originating in the bulk from the surface contribution by varying the applied DC field [18]. A very interesting quantum mechanical ab initio calculation of EFISH has been reported by Kikuchi and Tada [19]. Finally, it is worth mentioning that in the important contribution of Ohlhoff et al. [20] on EFISH at metal-oxide-semiconductor and metal-semiconductor structures, the difference between the s- and p-polarized EFISH responses is measured, but not explained. In that paper they compare experimental results with a model based on Green’s function due to Sipe [21] and also characterize the second harmonic intensity signal fitting Fourier coefficients for a series expansion.

In this work we argue based on group theoretical arguments why EFISH for light at normal incidence is not observed in all low index facets. However, for light with oblique incidence, a rather strong p-polarized EFISH and a rather weak s-polarized EFISH occur. A nonlinear bond model can be very illustrative [22,23] for explaining SHG. In a previous work, we sketched the main ideas for understanding the origin of EFISH using effective third-rank tensors and discussed the resulting symmetry after contracting it with the electric field in the z direction, establishing an analogy with the surface break of the symmetry [24]. Also there, we extend the so-called simplified bond hyperpolarizability model (SBHM) [25,26] for EFISH. In the present paper, we analyze the full fourth-rank tensor description with the symmetry breaking the perturbing electric field, presenting explicitly the 81 elements (many of them zero, as is well known) in terms of the contracted (engineering or Voigt) notation for silicon and zincblende having (001), (011), and (111) facets, for both the s- and p-polarization.

It is worth mentioning that in this model, we compute the pure EFISH bulk effect of perfect crystalline Si, which will be in any SHG measurement superimposed with the coherent contribution of surfaces, crystal defects, and grain boundaries in the case of polycrystalline Si, etc. To describe effects at defect or impurity sites breaking the Oh symmetry, group theory can be used in the context of crystal field theory (see, e.g., Ref. [27]), but this is beyond our aim of describing crystalline and defect free semiconductors.

The paper is organized as follows. Section 2 is a brief discussion of the “well-known” fourth-rank tensors or third-order susceptibility predicted by group theory for Si applied to the (001), (011), and (111) facets. The consequence of “symmetrizing” the fourth-rank tensor due to the presence of two identical electromagnetic fields is explained, reducing the number of independent elements to two. The resulting nonlinear polarizabilities (for incident p- and s-polarization) are calculated for these “symmetrized” tensors for each facet, and the resulting SH signal is provided. In Section 3, a classical oscillator model is used for a pictorial view of the interaction between the electric fields and the electrons in the crystal, pointing out the importance of the symmetry breaking through the direction of the oscillating electric field. Finally, conclusions are given in Section 4.

2. GROUP THEORETICAL DESCRIPTION OF EFISH FOR ZINCBLENDE AND DIAMOND (001), (111), AND (011) FACETS

The symmetries associated with a crystal can be described using matrices and, when connected to symmetries of space and time, by tensors. All the matrices describing the operations in a crystal form a group or, more specifically, a point group. The surfaces and interfaces are a natural break of the symmetry in a crystal, but this breaking can also be generated by other external factors. For example, in the nonlinear phenomenon known as EFISH generation, the break of symmetry is due to the application of a static electric field (DC electric field). Mathematically, for these two frequencies, ω and zero, the linear and nonlinear response of the system is modeled through the resulting polarization:

P(ω)=ϵ0{χ(1)·E(ω)+χ(2A)··[E(ω)E(0)]+}P(2ω)=ϵ0{χ(2)··[E(ω)E(ω)]+χ(3A)[E(ω)E(ω)E(0)]+}P(3ω)=ϵ0{χ(3)[E(ω)E(ω)E(ω)]+},
where P(ω) is the linear polarization and E(ω) the fundamental electric field; χ(1) is known as the first-order susceptibility, χ(2A) describes the linear electro-optic (sometimes called Kerr) effect (i.e., the modification of the dielectric function/refractive index through a DC field), χ(2) describes SHG, χ(3A) is the above mentioned EFISH generation, and finally, χ(3) represents THG. If the electric field is described in a full vectorial way, then, in general, χ(1), χ(2), and χ(3) are represented through second-, third- and fourth-rank tensors, respectively. We mention here that χ(3A) and χ(3) will not have the same values, but due to symmetry they have the same tensorial form.

In particular, the second term in Eq. (1) for P(2ω) precisely describes the correspondent third-order phenomena, which can be rewritten in terms of its components for the particular EFISH case by

Pi=χijkl(3)Ej(ω)Ek(ω)El(0).

As mentioned before, χijkl(3) is a fourth-rank tensor; the “i” index is the resulting component of the polarization. The indices “j” and “k” are contracted with the electric field, and these indices can be permuted, whereas the index “l” is contracted with the DC field; hence, it cannot be permuted with the other indices. A general fourth-rank tensor can be represented as a 3×3 matrix, which also has 3×3 matrices as elements, such as [28]

χ(3)=((χ1111χ1112χ1113χ1121χ1122χ1123χ1131χ1132χ1133)(χ1211χ1212χ1213χ1221χ1222χ1223χ1231χ1232χ1233)(χ1311χ1312χ1313χ1321χ1322χ1323χ1331χ1332χ1333)(χ2111χ2112χ2113χ2121χ2122χ2123χ2131χ2132χ2133)(χ2211χ2212χ2213χ2221χ2222χ2223χ2231χ2232χ2233)(χ2311χ2312χ2313χ2321χ2322χ2323χ2331χ2332χ2333)(χ3111χ3112χ3113χ3121χ3122χ3123χ3131χ3132χ3133)(χ3211χ3212χ3213χ3221χ3222χ3223χ3231χ3232χ3233)(χ3311χ3312χ3313χ3321χ3322χ3323χ3331χ3332χ3333)).

There are 81 elements, which are labeled as follows. The first index “i” in χijkl(3) corresponds to the rows and the second index “j” to the columns in the main matrix (the external one). In the same way, the indices “k” and “l” will correspond to the usual way of labeling a 3×3 matrix, namely, the rows and columns in the inner 3×3 matrix, respectively. For example, the element χ3213 corresponds to the element in the external matrix within position third row and second column but now this element is also a matrix and inside that matrix we should look for the element in the first row and third column.

On the other hand, there is a more compact way to label the elements in the tensor, known as the engineering or Voigt matrix representation. Using this representation, the tensor in Eq. (3) can be written in the form of a 6×6 matrix, given by

S=(s11s12s13s14s15s16s21s22s23s24s25s26s31s32s33s34s35s36s41s42s43s44s45s46s51s52s53s54s55s56s61s62s63s64s65s66),
where one usually assumes symmetry along the diagonal. This is for example the case of the compliance tensor (see for example Ref. [29]), connecting the symmetric stress and strain tensors, which has a lower number of independent elements. As is well known, the relationships between tensor subscripts for the contracted notation with two indices and the one with four indices are given by 111, 222, 333, 423or32, 531 or 13, and 612 or 21. Also, smn=χijkl when m and n are 1, 2 or 3; smn=2χijkl when either m or n is 4, 5, or 6; and smn=4χijkl when both m and n are 4, 5, or 6.

However, the transformation rule for tensors can only be applied under the form given by Eq. (3), which is the explicit fourth-rank tensor with four subscripts. The general transformation is then

χijkl=RimRjnRkoRlpχmnop,
where Rab is a matrix defining a symmetry operation or could be general rotation of an arbitrary angle φ (or mirror or inversion). Applying the matrix in the transformation is a symmetry operation belonging to the point group of the crystal; then the initial tensor and the final one in Eq. (5) must be the same (χijkl=χmnop) for physical reasons. This is also known as Neumanns’ principle [30]. This condition is the one that, according to the symmetry of the crystal, determines the number of nonzero independent elements in the susceptibility tensor. In the next subsections we will discuss the resulting symmetries and their tensors for the Si(001), Si(111), and Si(011) facets when a static electric field is applied along the z-axis, breaking in this way all the symmetry operations implying this Cartesian direction.

A. EFISH of Zincblende (001) Facet

It is well known that silicon (zincblende) crystals have Oh(Td) symmetry. However, when we study their surfaces, the symmetry is reduced, depending on the direction in which the crystal has been cut. Also, for the case of a static electric field applied in a particular direction, the symmetry is broken due to the perturbation, and the resulting group has less symmetry elements.

On the other hand, the response of the conventional cell can be very well simulated through SBHM only by using one tetrahedral element. A tetrahedron has Td symmetry, but when the symmetry is broken in one direction, as mentioned before, the symmetry is lowered. When a DC electric field is applied in the z direction in the coordinate system shown in Fig. 1, the electrons are shifted from their equilibrium position inside the bulk (in the middle between the atoms), and the bonds are no longer equivalent or equal. Thus, the symmetry is now reduced to C2v point group, with a two-fold axis of rotation around z [20].

 figure: Fig. 1.

Fig. 1. Conventional silicon cell.

Download Full Size | PDF

Mathematically, the tensor for an Oh as well as a Td point group can be stated in contracted notation, without any symmetry breaking through an external DC field [23]:

χGT(3)(001)=(s11s12s12000s21s11s12000s21s21s11000000s44000000s44000000s44),
where the parenthesis (001) is only to remember the orientation of the crystal, which also matches the direction in which the DC field is going to be applied, because later we are going to rotate this tensor to reproduce the orientations of the other facets. Therefore, in the four index representation, Eq. (6) changes to
χijkl,GT(3)(001)=((s11000s12000s12)(0s4440s44400000)(00s444000s44400)(0s4440s44400000)(s21000s11000s12)(00000s4440s4440)(00s444000s44400)(00000s4440s4440)(s21000s21000s11)),
where there are still four independent elements in the tensor. In the case of EFISH given in Eq. (2), the second and third indices in the tensor can be permuted because there is only one input frequency ω. When “symmetrizing” the tensor, it is possible to reduce the number of independent elements to two. This is because
χijkl,EFISH(3)(2ω,ω,ω,0)=12(χijkl,EFISH(3)+χikjl,EFISH(3)),
and then s444=s12=s21. We now contract the tensor in Eq. (7) with the electromagnetic fields. After contracting the susceptible fourth-rank tensor with the DC field, the resulting tensor is of the third-rank, and its symmetry changes from Oh to C2v. Thus, the physical considerations of the symmetry breaking agree with the mathematical results. Additionally, for s-polarization, the electric field will only have components in the y-direction Es=(0,Ey(ω),0), whereas the DC electric field always will be applied in the z direction EDC=(0,0,Ez(0)). Thus, neglecting for shortness the Fresnel coefficients, which are necessary to calculate the measured output electric field produced by bulk radiations, the nonlinear induced polarization can be written as follows [independent of the angle of incidence; see Fig. 2(a)]:
Ps(2ω)=(0,0,s12Ey2Ez).

 figure: Fig. 2.

Fig. 2. (a) SHG, inside the high index nonlinear material, has a very small angle with the normal. There is a coherent contribution just for a certain depth d. (b) High index prims keeps a large angle with the normal inside the nonlinear material.

Download Full Size | PDF

This radiation, however, has only a z-component; that is, it is equivalent that the polarization direction for the frequency doubled light of the dipoles is along z^. The resulting radiation can therefore not be detected in normal incidence transmission or reflection SHG, because in the ambient medium (air or vacuum), the polarization of the displacement field has to be normal to the propagation direction. Under oblique incidence, however, the s-polarized EFISH signal should become weakly measurable in transmission (because refraction is towards the surface normal) and to an even weaker extent in reflection.

The notation for the angle of incidence and refraction and the second harmonic angles is depicted in Fig. 2(a). Due to Snell’s law, the refracted beam always propagates rather close to the surface normal in silicon; so, the radiated intensity proportional to |P|2 will be rather weak and will only be detected for a p-polarization, which will be discussed now.

For a p-polarization, the electric field will only have components in the x and z directions Ep=Ep(cosθ2(ω),0,sinθ2(ω)), and the DC field is again along z. The result is

Pp(2ω)=Ep2Ez(s12sin2θ2(ω),0,s11sin2θ2(ω)+s12cos2θ2(ω)),
where θ2(ω) is the angle of transmission after applying Snell’s law. However, in materials with a rather high refractive index, Snell’s law will reduce the angle of incidence considerably and thereby diminish the EFISH effect. Also note that Eq. (10) mainly has a z-component. Using a high index prism for incoupling to keep θ2(ω) large could enhance the effect [see Fig. 2(b)], and in this way, this signal could be measurable. Here it is worth mentioning that the direction and amplitude of the electric fields in air are known. Inside the Si material, the fields can be calculated using Fresnel coefficients and their directions by Snell’s law [31] as follows:
n1(ω)sinθ1(ω)=n2(ω)sinθ2(ω),
whereas the same is valid for the SH signal:
n2(2ω)sinθ2(2ω)=n1(2ω)sinθ1(2ω).

It is worth noticing that, in general, ni(ω)ni(2ω) and θi(ω)θi(2ω) (i=1, 2). Additionally, for optically dense materials, the propagation direction approaches the normal, and there is a very small contribution for the x-component in the p-polarization when the incident angle is small. Therefore, for normal incidence, there is no measurable EFISH response for the s-polarization, and it will only show up for the p-polarization, whereas Eq. (10) reduces to Eq. (9) as should be under normal incidence.

B. EFISH of Zincblende (111) Facet

As the orientation is different, we need to rotate the crystal to a system of reference with the z-axis perpendicular to the plane (111). To do this, two rotations must be performed; the first one is around the z-axis by an angle of π/4; then in the new system of reference, the second rotation will be around the y-axis by β/2, where β is the angle between the bonds, which it is equal to 109.47° for silicon. Therefore, after applying Eq. (5) to the tensor given in Eq. (7) and taking account of the “symmetrization” in indices “j” and “k,” the resulting tensor is

χijkl,EFISH(3)(111)=((12(s11+3s12)0s113s1232016(s11+3s12)0s113s12320s113)(016(s11+3s12)016(s11+3s12)0s113s12320s113s12320)(s113s12320s1130s113s12320s11300)(016(s11+3s12)016(s11+3s12)0s113s12320s113s12320)(16(s11+3s12)0s113s1232012(s11+3s12)0s113s12320s113)(0s113s12320s113s12320s1130s1130)(s113s12320s1130s113s12320s11300)(0s113s12320s113s12320s1130s1130)(s113000s11300013(s11+6s12))).

This tensor still has Oh symmetry, but in the new system of reference it looks rather different. Two independent elements remain, but now there are five different values in the tensor. Although, when this fourth-rank tensor Eq. (13) is contracted with the DC field, the resulting symmetry group for the third-rank tensor is C3v (the three-fold axis of rotation is the z direction), as it should be. This can be checked directly doing the contraction, as mentioned before, or for physical arguments when the symmetry operations involving the z coordinate are eliminated.

As before, we can contract the tensor in Eq. (13) with all the electromagnetic fields defined in the same way and get the final expressions for the incident polarizations. For the case of s-polarization,

Ps(2ω)=13Ey2Ez(22[s113s12],0,s11),
whereas the p-polarization yields
Pp(2ω)=13Ep2Ez(22[s113s12]cos2θ2(ω)s11sin2θ2(ω),0,s11+6s12sin2θ2(ω)),
which is more complicated due to the presence of the two components in the incident electromagnetic field. For normal incidence, the dipoles now oscillate in the xz plane, and this radiation (more precise its x-component) can be coupled out in transmission and reflection. To determine the absolute values of the effect, the Fresnel transmittances for 2ω should be taken into account. A possible way to establish a rate between components s11 and s12 is through Brewster’s angle relation for the electric fields at the 2ω frequency; in this case, the electric fields oscillating at second harmonic propagate and have the linear description
P(2ω)=ϵ0χ(2ω)E(2ω),
where the susceptibility χ(2ω) is now an scalar.

C. EFISH of Zincblende (110) Facet

For the same reasons exposed in the previous section, here we are going to start with the tensor given in Eq. (7) after the symmetrization in the indices “j” and “k.” However, the rotations will be first π/4 around the z-axis and then π/2 around the y-axis. These two combined rotations result in the following tensor:

χijkl,EFISH(3)(110)=((s11000s12000s12)(0s120s1200000)(00s12000s1200)(0s120s1200000)(s12000s11+3s122000s11s122)(00000s11s1220s11s1220)(00s12000s1200)(00000s11s1220s11s1220)(s12000s11s122000s11+3s122)).

Please note that even when the number of independent coefficients in the tensor is the same, there are four different values in it. The contraction of this tensor Eq. (17) with the DC field and the electromagnetic fields results in

Ps(2ω)=12Ey2Ez(0,0,[s11s12]),
for the s-polarization. Here Ps(2ω)=0 if Kleinman symmetry holds (s11=s12), which is predicted by SBHM; then s-polarization should not contribute to the SHG signal, whereas the p-polarization yields
Pp(2ω)=12Ep2Ez(2s12sin2θ2(ω),0,2s12+[s11+s12]sin2θ2(ω)).

In analogy with Eq. (15), this expression for Pp(2ω) can be used as a test for relative magnitudes via Brewster’s angle. It is interesting that in this case, if θ1(ω)=0, a different result is obtained than the one for s-polarization, whereas in the previous directions Si(001) and Si(111), they reduce to the same, just changing EyEp. This means that for the Si(110) facet, the directions x and y are not fully equivalent, a fact that is already known from linear optics. In linear optics, the polarization sensitive technique reflection difference/anisotropy spectroscopy can distinguish crystal orientations for the case of Si(110) facets [32].

As a summary of this section, we would like to point out that for Si(001) and Si(110) no EFISH signal should be detected for normal incidence, neither in transmission nor in reflection, because the second harmonic electric field only has a component for the z direction. In a Si(111) facet, a strong EFISH signal for both the s- and p-polarizations should be detected due to the breaking of symmetry in the bulk for this bond configuration. In addition, we point out that the electric field generated through the EFISH effect, which is propagating along the z direction, could interfere destructively at the detector if the region where the static electric field is applied is longer than the coherence length between the fundamental and harmonic. However, for usual doping concentrations and semiconductor depletion regions, this will hardly occur in transmission, but it could be significant in reflection.

In the next section, we will present a phenomenological explanation of EFISH using a bond model, where the physics of this phenomenon is very clear by considering the movement of electrons along the bonds due to the action of the electromagnetic fields.

3. CLASSICAL MODEL FOR EFISH

There is no dipole-related SHG signal in silicon and, in general, in any centrosymmetric crystal due to bulk contribution, at least attributed to a second-order nonlinear phenomenon. Still, there could be contributions from several other mechanisms, for example, built-in strain [22], but we are only going to discuss EFISH here, where the breaking of the symmetry in the electronic distribution originates from a static electric field applied in the z direction.

Now, we would like to pictorially explain EFISH mechanism, arguing about the break of the symmetry in a conventional silicon cell and in the physics of the modifications suffered by the electric charge density along the bonds when the DC electric field is applied.

As shown mathematically before, there is no EFISH signal for normal incidence (s-polarization) for facets Si(001) and Si(110). This is because even when breaking the symmetry (in the z direction because of the DC field), there are always bonds opposing each other [see Figs. 3(a) and 3(b)], and canceling in this way, the signal contribution is generated by the incoming radiation, with only one component in its electromagnetic field. So, even though the symmetry is broken by the static field along z, for normal incidence, as seen in Figs. 3(a) and 3(b), the middle blue silicon atom forms a two-dimensional in plane (xy) inversion center.

 figure: Fig. 3.

Fig. 3. Equilibrium position of the electric charge is shifted due to the DC field EZ. Facets (a) Si(001), (b) Si(110), and (c) Si(111).

Download Full Size | PDF

However, under oblique incidence this statement is not fully true anymore: p-polarization has two components in the electromagnetic field, and therefore, due to the interaction of the two orthogonal components driving the electrons, the opposing bonds no longer equilibrate each other, resulting in an EFISH signal.

In the case of the facet Si(111), the DC field also breaks the symmetry in the direction z, but here due to the orientation, there is no opposing bond that represents the response of the conventional cell [see Fig. 3(c), where down bonds are below the xy plane and one is oriented directly below the negative y-axis]. In fact, the bond parallel to the z-axis provides no contribution at all for EFISH under normal incidence, because for normal incidence this wave function has still a 2D (xy) inversion symmetry, whereas the other three bonds pointing down have not (a 2D xy inversion symmetry) and contribute to EFISH.

For the three bonds, which are tilted with respect to the static electric field, we now define their equation of motion by taking into account the static field potential: classically in terms of the atomic potential only containing even powers [33]. Thus, the potential for each tilted bond material is

V(r)=12mω02r214mbr4qsinγEDCr,
where m is the mass of the electron, ω0 is the oscillator resonance frequency, b is a parameter that characterizes the strength of the nonlinearity (THG in this case), and r is the radial coordinate. The last term corresponds to the applied DC electric field, where q is the electron charge, γ is the angle between the down bonds and the xy plane [see Fig. 3(c)], and, of course, EDC is the applied static electric field. This originates a break in the atomic potential symmetry, which consists of a shift from the equilibrium position in the electronic distribution. The atomic lattice can be seen as fixed in the process, and therefore, the electronic shift can be described as a break in the original symmetry configuration. As the nonlinear effect is a small amount of the incident original signal, we assume br4ω0r2 and look for the new equilibrium position in the force equation, then
rr0qsinγEDCmω02=r0reff,
where r0 is the original equilibrium position (in the middle between two nuclei) before the break of the symmetry. Expanding the potential V(r) around this new equilibrium position yields
V(r)=12(mω02+12mbreff2)reff2mbreff3(rreff)+12(mω023mbreff2)(rreff)2mbreff(rreff)3+O[rreff]4,
where reff=qsinγEDCmω02. The first term in Eq. (23) gives the constant, which changes the potential when applying a static electric field, but most importantly, the potential now comprises a term proportional to (rreff)3, which produces, via the equation of motion, SHG in the force equation. The SHG-EFISH hyperpolarizability is given by
aeff(2)(2ω)=3bqω02sinγEDC.

The interesting result from Eq. (22) is that the strength parameter b for THG will be related to the second harmonic signal originated through EFISH mechanism.

Remarkably, Kikuchi and Tada [19] have already shown, using quantum mechanical calculations, that EFISH can be modeled with only one independent parameter in the fourth-rank tensor if adequate conditions exist. Therefore, this result is very reasonable and agrees with experimental results for SHG [34,35] and EFISH [20].

4. CONCLUSIONS

The EFISH generation is clearly a bulk phenomenon. Whereas group theory requires only two independent elements in the fourth-rank susceptibility tensor (after symmetrization), the classical oscillator bond model fits the experimental data reasonably well with only one free parameter. Moreover, both approaches predict an EFISH signal for facet Si(111) but none for facets Si(110) and Si(001) for normal incidence. We also propose experimental setups, which can test the validity of SBHM for EFISH generation. In general, these results are valid for zincblende crystals, but it could be difficult to separate the dipole allowed contribution from the EFISH effect. However, we think that an appropriated combination of the angle and polarization dependence should enable one to distinguish between these two phenomena.

The existence of a relationship between the parameter that characterizes THG and the second harmonic signal originated through EFISH is a very interesting result, and the validity should be checked by experiments or ab initio theories.

Funding

“MAYANET” Erasmus Mundus Partnership (EU-2015-578); Consejo Nacional de Ciencia y Tecnología (CONACYT) (CB-2015-01/254617); Horizon 2020 Framework Programme (H2020) (GA692034).

Acknowledgment

A. A. M. thanks the program “MAYANET” Erasmus Mundus Partnership—Action 2 for the mobility scholarship level “TG 2 Academic Staff” and CONACyT for financial support during this research (“Theoretical Description of Nonlinear Optical Interactions for Surfaces and Bulk, Using Simplified Bond Hyperpolarizability Model and Group Theory”). K. H. acknowledges the support of the European Commission under the H2020 grant TWINFUSYON.

REFERENCES

1. J. P. Gordon, H. J. Zeiger, and C. H. Townes, “The maser, new type of microwave amplifier, frequency, standard and spectrometer,” Phys. Rev. 99, 1264–1274 (1955). [CrossRef]  

2. P. A. Franken, A. E. Hill, C. W. Peters, and G. Weinreich, “Generation of optical harmonics,” Phys. Rev. Lett. 7, 118–119 (1961). [CrossRef]  

3. P. D. Maker and R. W. Terhune, “Study of optical effects due to an induced polarization third order in the electric field strength,” Phys. Rev. 137, A801–A818 (1965). [CrossRef]  

4. J. F. Ward and G. H. C. New, “Optical third harmonic generation in gases by a focused laser beam,” Phys. Rev. 185, 57–72 (1969). [CrossRef]  

5. R. W. Terhune, P. D. Maker, and C. M. Savage, “Optical harmonic generation in calcite,” Phys. Rev. Lett. 8, 404–406 (1962). [CrossRef]  

6. Y.-S. Lee, M. H. Anderson, and M. C. Downer, “Fourth-harmonic generation at a crystalline GaAs(001) surface,” Opt. Lett. 22, 973–975 (1997). [CrossRef]  

7. Y.-S. Lee and M. C. Downer, “Reflected fourth-harmonic radiation from a centrosymmetric crystal,” Opt. Lett. 23, 918–920 (1998). [CrossRef]  

8. Y.-S. Lee and M. C. Downer, “Reflected optical fourth harmonic generation at crystalline surfaces,” Thin Solid Films 364, 80–85 (2000). [CrossRef]  

9. J.-K. Hansen, H. J. Peng, and D. E. Aspnes, “Application of the simplified bond-hyperpolarizability model to fourth-harmonic generation,” J. Vac. Sci. Technol. B 21, 1798–1803 (2003). [CrossRef]  

10. J. A. Giordmaine and R. C. Miller, “Tunable coherent parametric oscillation in LiNbO3 at optical frequencies,” Phys. Rev. Lett. 14, 973–976 (1965). [CrossRef]  

11. J. A. Giordmaine and R. C. Miller, “Optical parametric oscillation in the visible spectrum,” Appl. Phys. Lett. 9, 298–300 (1966). [CrossRef]  

12. R. W. Boyd and C. H. Townes, “An infrared upconverter for astronomical imaging,” Appl. Phys. Lett. 31, 440–442 (1977). [CrossRef]  

13. R. L. Byer and R. L. Herbst, Tunable Infrared Generation, Y. R. Shen, ed. (Springer, 1977).

14. C. H. Lee, R. K. Chang, and N. Bloembergen, “Nonlinear electroreflectance in silicon and silver,” Phys. Rev. Lett. 18, 167–170 (1967). [CrossRef]  

15. S. Kielich, “DC electric field-induced second harmonic light generation in gases and liquids,” Acta Phys. Pol. A37, 205–219 (1970).

16. C. G. Bethea, “Electric field induced second harmonic generation in glass,” Appl. Opt. 14, 2435–2437 (1975). [CrossRef]  

17. O. A. Aktsipetrov, A. A. Fedyanin, V. N. Golovkina, and T. V. Murzina, “Optical second-harmonic generation induced by a dc electric field at the Si-SiO2 interface,” Opt. Lett. 19, 1450–1452 (1994). [CrossRef]  

18. O. A. Aktsipetrov, A. A. Fedyanin, and A. V. Melnikov, “DC electric field induced second-harmonic generation spectroscopy of the Si(001)-SiO2 interface: separation of the bulk and surface non-linear contributions,” Thin Solid Films 294, 231–234 (1997). [CrossRef]  

19. K. Kikuchi and K. Tada, “Theory of electric field-induced optical second harmonic generation in semiconductors,” Opt. Quantum Electron. 12, 199–205 (1980). [CrossRef]  

20. C. Ohlhoff, G. Lüpke, C. Meyer, and H. Kurz, “Static and high-frequency electric fields in silicon MOS and MS structures probed by optical second-harmonic generation,” Phys. Rev. B 55, 4596–4606 (1997). [CrossRef]  

21. J. E. Sipe, “New Green-function formalism for surface optics,” J. Opt. Soc. Am. B 4, 481–489 (1987). [CrossRef]  

22. D. E. Aspnes, “Bond models in linear and nonlinear optics,” Phys. Status Solidi B 247, 1873–1880 (2010). [CrossRef]  

23. D. E. Aspnes, “Bond models in linear and nonlinear optics,” Proc. SPIE 9584, 95840A (2015). [CrossRef]  

24. A. Alejo-Molina, K. Hingerl, and H. Hardhienata, “Model of third harmonic generation and electric field induced optical second harmonic using simplified bond-hyperpolarizability model,” J. Opt. Soc. Am. B 32, 562–570 (2015). [CrossRef]  

25. G. D. Powell, J. F. Wang, and D. E. Aspnes, “Simplified bond hyperpolarizability model of second harmonic generation,” Phys. Rev. B 65, 205320 (2002). [CrossRef]  

26. J.-F. T. Wang, G. D. Powell, R. S. Johnson, G. Lucovsky, and D. E. Aspnes, “Simplified bond-hyperpolarizability model of second harmonic generation: application to Si-dielectric interfaces,” J. Vac. Sci. Technol. B 20, 1699–1705 (2002). [CrossRef]  

27. E. Pavarini, “Crystal-field theory, tight-binding method, and Jahn-Teller effect,” in Correlated Electrons: From Models to Materials, E. Koch, F. Anders, and M. Jarrell, eds. (Forschungszentrum Jülich, 2012), pp. 6.1–6.39.

28. R. C. Powell, Symmetry, Group Theory, and the Physical Properties of Crystals, Lecture Notes in Physics (Springer, 2010), Vol. 824.

29. P. Y. Yu and M. Cardona, Fundamental of Semiconductors, Physics and Materials Properties (Springer, 2010).

30. J. F. Nye, Physical Properties of Crystals, Their Representations by Tensors and Matrices (Clarendon, 1957).

31. J. D. Jackson, Classical Electrodynamics, 3rd ed. (Wiley, 1998).

32. D. E. Aspnes and A. A. Studna, “Anisotropies in the above—band-gap optical spectra of cubic semiconductors,” Phys. Rev. Lett. 54, 1956–1959 (1985). [CrossRef]  

33. R. W. Boyd, Nonlinear Optics, 2nd ed. (Academic, 2003).

34. H. Hardhienata, A. Prylepa, D. Stifter, and K. Hingerl, “Simplified bond-hyperpolarizability model of second-harmonic generation in Si(111): theory and experiment,” J. Phys. 423, 012046 (2013). [CrossRef]  

35. H. Hardhienata, A. Alejo-Molina, C. Reitböck, A. Prylepa, D. Stifter, and K. Hingerl, “Bulk dipolar contribution to second-harmonic generation in zincblende,” J. Opt. Soc. Am. B 33, 195–201 (2016). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (3)

Fig. 1.
Fig. 1. Conventional silicon cell.
Fig. 2.
Fig. 2. (a) SHG, inside the high index nonlinear material, has a very small angle with the normal. There is a coherent contribution just for a certain depth d . (b) High index prims keeps a large angle with the normal inside the nonlinear material.
Fig. 3.
Fig. 3. Equilibrium position of the electric charge is shifted due to the DC field E Z . Facets (a) Si(001), (b) Si(110), and (c) Si(111).

Equations (23)

Equations on this page are rendered with MathJax. Learn more.

P ( ω ) = ϵ 0 { χ ( 1 ) · E ( ω ) + χ ( 2 A ) · · [ E ( ω ) E ( 0 ) ] + } P ( 2 ω ) = ϵ 0 { χ ( 2 ) · · [ E ( ω ) E ( ω ) ] + χ ( 3 A ) [ E ( ω ) E ( ω ) E ( 0 ) ] + } P ( 3 ω ) = ϵ 0 { χ ( 3 ) [ E ( ω ) E ( ω ) E ( ω ) ] + } ,
P i = χ i j k l ( 3 ) E j ( ω ) E k ( ω ) E l ( 0 ) .
χ ( 3 ) = ( ( χ 1111 χ 1112 χ 1113 χ 1121 χ 1122 χ 1123 χ 1131 χ 1132 χ 1133 ) ( χ 1211 χ 1212 χ 1213 χ 1221 χ 1222 χ 1223 χ 1231 χ 1232 χ 1233 ) ( χ 1311 χ 1312 χ 1313 χ 1321 χ 1322 χ 1323 χ 1331 χ 1332 χ 1333 ) ( χ 2111 χ 2112 χ 2113 χ 2121 χ 2122 χ 2123 χ 2131 χ 2132 χ 2133 ) ( χ 2211 χ 2212 χ 2213 χ 2221 χ 2222 χ 2223 χ 2231 χ 2232 χ 2233 ) ( χ 2311 χ 2312 χ 2313 χ 2321 χ 2322 χ 2323 χ 2331 χ 2332 χ 2333 ) ( χ 3111 χ 3112 χ 3113 χ 3121 χ 3122 χ 3123 χ 3131 χ 3132 χ 3133 ) ( χ 3211 χ 3212 χ 3213 χ 3221 χ 3222 χ 3223 χ 3231 χ 3232 χ 3233 ) ( χ 3311 χ 3312 χ 3313 χ 3321 χ 3322 χ 3323 χ 3331 χ 3332 χ 3333 ) ) .
S = ( s 11 s 12 s 13 s 14 s 15 s 16 s 21 s 22 s 23 s 24 s 25 s 26 s 31 s 32 s 33 s 34 s 35 s 36 s 41 s 42 s 43 s 44 s 45 s 46 s 51 s 52 s 53 s 54 s 55 s 56 s 61 s 62 s 63 s 64 s 65 s 66 ) ,
χ i j k l = R i m R j n R k o R l p χ m n o p ,
χ G T ( 3 ) ( 001 ) = ( s 11 s 12 s 12 0 0 0 s 21 s 11 s 12 0 0 0 s 21 s 21 s 11 0 0 0 0 0 0 s 44 0 0 0 0 0 0 s 44 0 0 0 0 0 0 s 44 ) ,
χ i j k l , G T ( 3 ) ( 001 ) = ( ( s 11 0 0 0 s 12 0 0 0 s 12 ) ( 0 s 44 4 0 s 44 4 0 0 0 0 0 ) ( 0 0 s 44 4 0 0 0 s 44 4 0 0 ) ( 0 s 44 4 0 s 44 4 0 0 0 0 0 ) ( s 21 0 0 0 s 11 0 0 0 s 12 ) ( 0 0 0 0 0 s 44 4 0 s 44 4 0 ) ( 0 0 s 44 4 0 0 0 s 44 4 0 0 ) ( 0 0 0 0 0 s 44 4 0 s 44 4 0 ) ( s 21 0 0 0 s 21 0 0 0 s 11 ) ) ,
χ i j k l , EFISH ( 3 ) ( 2 ω , ω , ω , 0 ) = 1 2 ( χ i j k l , EFISH ( 3 ) + χ i k j l , EFISH ( 3 ) ) ,
P s ( 2 ω ) = ( 0,0 , s 12 E y 2 E z ) .
P p ( 2 ω ) = E p 2 E z ( s 12 sin 2 θ 2 ( ω ) , 0 , s 11 sin 2 θ 2 ( ω ) + s 12 cos 2 θ 2 ( ω ) ) ,
n 1 ( ω ) sin θ 1 ( ω ) = n 2 ( ω ) sin θ 2 ( ω ) ,
n 2 ( 2 ω ) sin θ 2 ( 2 ω ) = n 1 ( 2 ω ) sin θ 1 ( 2 ω ) .
χ i j k l , EFISH ( 3 ) ( 111 ) = ( ( 1 2 ( s 11 + 3 s 12 ) 0 s 11 3 s 12 3 2 0 1 6 ( s 11 + 3 s 12 ) 0 s 11 3 s 12 3 2 0 s 11 3 ) ( 0 1 6 ( s 11 + 3 s 12 ) 0 1 6 ( s 11 + 3 s 12 ) 0 s 11 3 s 12 3 2 0 s 11 3 s 12 3 2 0 ) ( s 11 3 s 12 3 2 0 s 11 3 0 s 11 3 s 12 3 2 0 s 11 3 0 0 ) ( 0 1 6 ( s 11 + 3 s 12 ) 0 1 6 ( s 11 + 3 s 12 ) 0 s 11 3 s 12 3 2 0 s 11 3 s 12 3 2 0 ) ( 1 6 ( s 11 + 3 s 12 ) 0 s 11 3 s 12 3 2 0 1 2 ( s 11 + 3 s 12 ) 0 s 11 3 s 12 3 2 0 s 11 3 ) ( 0 s 11 3 s 12 3 2 0 s 11 3 s 12 3 2 0 s 11 3 0 s 11 3 0 ) ( s 11 3 s 12 3 2 0 s 11 3 0 s 11 3 s 12 3 2 0 s 11 3 0 0 ) ( 0 s 11 3 s 12 3 2 0 s 11 3 s 12 3 2 0 s 11 3 0 s 11 3 0 ) ( s 11 3 0 0 0 s 11 3 0 0 0 1 3 ( s 11 + 6 s 12 ) ) ) .
P s ( 2 ω ) = 1 3 E y 2 E z ( 2 2 [ s 11 3 s 12 ] , 0 , s 11 ) ,
P p ( 2 ω ) = 1 3 E p 2 E z ( 2 2 [ s 11 3 s 12 ] cos 2 θ 2 ( ω ) s 11 sin 2 θ 2 ( ω ) , 0 , s 11 + 6 s 12 sin 2 θ 2 ( ω ) ) ,
P ( 2 ω ) = ϵ 0 χ ( 2 ω ) E ( 2 ω ) ,
χ i j k l , EFISH ( 3 ) ( 110 ) = ( ( s 11 0 0 0 s 12 0 0 0 s 12 ) ( 0 s 12 0 s 12 0 0 0 0 0 ) ( 0 0 s 12 0 0 0 s 12 0 0 ) ( 0 s 12 0 s 12 0 0 0 0 0 ) ( s 12 0 0 0 s 11 + 3 s 12 2 0 0 0 s 11 s 12 2 ) ( 0 0 0 0 0 s 11 s 12 2 0 s 11 s 12 2 0 ) ( 0 0 s 12 0 0 0 s 12 0 0 ) ( 0 0 0 0 0 s 11 s 12 2 0 s 11 s 12 2 0 ) ( s 12 0 0 0 s 11 s 12 2 0 0 0 s 11 + 3 s 12 2 ) ) .
P s ( 2 ω ) = 1 2 E y 2 E z ( 0,0 , [ s 11 s 12 ] ) ,
P p ( 2 ω ) = 1 2 E p 2 E z ( 2 s 12 sin 2 θ 2 ( ω ) , 0,2 s 12 + [ s 11 + s 12 ] sin 2 θ 2 ( ω ) ) .
V ( r ) = 1 2 m ω 0 2 r 2 1 4 m b r 4 q sin γ E DC r ,
r r 0 q sin γ E DC m ω 0 2 = r 0 r eff ,
V ( r ) = 1 2 ( m ω 0 2 + 1 2 m b r eff 2 ) r eff 2 m b r eff 3 ( r r eff ) + 1 2 ( m ω 0 2 3 mb r eff 2 ) ( r r eff ) 2 m b r eff ( r r eff ) 3 + O [ r r eff ] 4 ,
a eff ( 2 ) ( 2 ω ) = 3 b q ω 0 2 sin γ E DC .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.