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

Terahertz-induced cascaded interactions between spectra offset by large frequencies

Open Access Open Access

Abstract

We explore the dynamics of a system where input spectra in the optical domain with very disparate center frequencies are strongly coupled via highly phase-matched, cascaded second-order nonlinear processes driven by terahertz radiation. The only requirement is that one of the input spectra contain sufficient bandwidth to generate the phase-matched terahertz-frequency driver. The frequency separation between the input spectra (or pump and seed spectra) can be more than ten times larger than the phase-matched terahertz frequency. This is in contrast to our previous work on cascaded parametric amplification, where the frequency separation between the pump and seed is required to be equal to the phase-matched terahertz frequency. A practical application of such a system where the cascading of a narrowband pump line centered at 1064 nm induced by a group of weaker seed lines centered about 1030 nm and separated by the phase-matched terahertz frequency is introduced. This approach is predicted to generate terahertz radiation with percent-level conversion efficiencies and millijoule-level pulse energies in cryogenically-cooled periodically poled lithium niobate. A model that solves for the nonlinear coupled interaction of terahertz and optical waves is employed. The calculations account for second and third-order nonlinearities, dispersion in the optical and terahertz domains as well as terahertz absorption. Ramifications of pulse formats on laser-induced damage are estimated by tracking the generated free-electron density. Strategies to mitigate laser-induced damage are outlined.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Nonlinear optical interactions which give rise to further coupled nonlinear interactions in a cascade of processes maybe broadly referred to as cascaded optical nonlinearities. A few examples are found in [1–6].

In many cases, cascaded nonlinearities are not all, well phase matched [2, 6]. However, this is not the case for difference-frequency generation between near-degenerate waves in the optical or near-infrared (NIR) spectral range yielding terahertz radiation. This is because terahertz frequencies are much smaller than optical frequencies, which enables many phase-matched processes to be simultaneously supported. Therefore, very interesting spectral dynamics are observed in such systems.

Cascading effects in terahertz generation systems have predominantly received attention in the context of optical-to-terahertz conversion efficiencies beyond the single-photon conversion or Manley-Rowe limit [3, 7–9]. In such cases, cascaded processes result in the repeated energy down-conversion of a single optical (or NIR) photon to produce multiple terahertz photons [3].

However, cascading effects in terahertz generation systems can be important even when absolute optical-to-terahertz conversion efficiencies are low. The simplest example of such a system is one with conventional optical parametric amplifier (OPA) like initial conditions. This involves a pump in the optical or NIR region and a signal/idler in the terahertz region. Several experiments in various configurations yielding low-to-modest conversion efficiencies along with modulated optical spectra have been reported in this regard [10–15].

It was theoretically shown that even for 0% optical-to-terahertz energy conversion efficiency, such configurations can only be described by considering cascading effects, i.e. the simultaneous evolution of a large number of coupled three-wave mixing processes [16].

 figure: Fig. 1

Fig. 1 Example of a previously studied terahertz generation system where it is necessary to consider the simultaneous evolution of many phase-matched second-order processes even if optical-to-terahertz energy conversion efficiencies are low. (a) Initial interactions in a cascaded terahertz parametric amplifier. An OPA-like input of a strong optical pump at ω0 and weaker optical seed at ω−1 is used. The frequency separation between pump and seed is required to be equal to the phase-matched terahertz frequency, i.e. ω0ω1 = ΩTHz. The double arrows delineate the pair of waves involved in the interaction and the dotted arrows indicate the product of those interactions. For an input comprised of a strong pump at ω0 and weak seed at ω1 in the optical domain, a terahertz wave at ΩTHz corresponding to the beat frequency is first generated which then drives via sum frequency generation, the component at ω+1. (b) In subsequent steps, the generated terahertz radiation drives generation of red-shifted components ω2,3 and further blue-shifted components, e.g. ω+2. SFG consumes a terahertz photon while DFG generates one. Therefore, absolute conversion efficiencies are low in the absence of a mechanism which prefers DFG.

Download Full Size | PDF

The reasons for this are apparent by inspecting the schematic in Fig. 1, which describes the spectral dynamics of the terahertz generation system with conventional OPA-like initial conditions studied in [16]. Consider the initial condition when a strong optical pump with angular frequency ω0 and optical seed of weaker intensity with angular frequency ω1=ω0ΩTHz are phase matched (i.e. Δk=0) for the difference-frequency generation (DFG) of terahertz radiation with angular frequency ΩTHz. It is worth pointing out that the frequency separation between the pump and seed is exactly equal to the phase-matched terahertz frequency. This is in contrast to the approach presented in this work, where the pump and seed may be separated by a frequency very different from the phase-matched terahertz frequency.

The initial interactions are depicted in Fig. 1(a). Here, the line at ω0 beats with ω1 to generate ΩTHz, which in turn beats with ω0 to generate ω1. Mathematically, the phase-mismatch Δk=k(ω0)k(ω0ΩTHz)k(ΩTHz)=0 (where k is the wave number). However, since ΩTHzω0, k(ω0+ΩTHz)k(ω0)k(ΩTHz) is also approximately zero. Therefore, the sum-frequency generation (SFG)of ω+1=ω0+ΩTHz is also phase matched. Each side-band ω±1 can undergo further phase-matched SFG and DFG processes to produce a series of lines separated by ΩTHz as shown in Fig. 1(b). Each SFG process consumes a terahertz photon while each DFG process creates one. Therefore, in the absence of a mechanism to preferentially phase-match DFG processes, only modulation of terahertz radiation will occur. At the same time, the optical spectrum would exhibit many side-bands [11, 14, 15]. This illustrates the importance of considering the simultaneous evolution of many strongly coupled nonlinear interactions or cascading effects in a system with OPA-like initial conditions.

The consideration of cascading effects for OPA-like inputs predict other anomalies [16]. For instance, the wave of highest intensity could be present at a lower frequency (ω1) to produce terahertz amplification, which is disallowed in conventional parametric amplifiers. This is because the input optical spectrum is rapidly washed out due to the simultaneous evolution of many strongly coupled phase-matched processes and hence becomes inconsequential.

Building on this understanding of the nature of strongly coupled phase-matched processes in terahertz generation systems, in the present work we explore the spectral dynamics of a much more sophisticated system. It is comprised of optical/NIR pump and seed spectra separated by a very large frequency offset. However, one of the input spectra must contain sufficient bandwidth to generate the phase-matched terahertz frequency. The generated terahertz radiation then drives cascaded interactions between the pump and seed spectra (See Section 2, Fig. 2 for an illustration). For example, the frequency separation between the pump and seed spectra maybe 10 THz. However, if the seed has a bandwidth >0.5 THz, then phase-matched generation of 0.5 THz can mediate the flow of energy from pump to seed, despite their separation being 10 THz. This is in contrast to the case in Fig. 1 and [15, 16], where the frequency separation between pump and seed is exactly equal to the phase-matched terahertz frequency.

The above format lends itself to useful practical applications. The first involves coupling a single narrowband line of high energy at 1064 nm to a set of much weaker seed lines centered at 1030 nm. The seed lines are separated by a phase-matched sub-terahertz frequency. It is worthwhile to note that the separation between pump and seed is about 10 THz, which is more than ten times larger than the generated terahertz frequency. Furthermore, note that the higher energy pump line is situated at a lower frequency in relation to the lower energy seed lines.

The terahertz radiation generated by beating of lines centered about 1030 nm, drives the cascading of the line at 1064 nm. Eventually, the two spectra merge and subsequently an effective red-shift of the total optical spectrum results to yield very efficient terahertz generation at the percent-level in cryogenically-cooled periodically poled lithium niobate (PPLN). The availability of high energy Nd:YAG lasers at 1064 nm makes this approach particularly relevant to high energy terahertz generation. For pump fluences on the order of 0.5 Jcm 2 in conjunction with cm 2-aperture crystals [17], these conversion efficiencies translate to millijoule-level terahertz pulse energies with hundreds of cycles. Such pulses are of interest for compact particle acceleration [18–20] and X-ray free-electron lasers [21].

In section 2, we outline the general physics of the approach. In section 3, we present detailed numerical studies describing the physics of the system. The effect of the input spectrum on laser-induced damage is studied and methods to mitigate it are suggested. In section 4, we present numerical results of terahertz generation using a 1064 nm pump and a set of seed lines centered at 1030 nm. We conclude in section 5.

2. Physical mechanism

In Fig. 2, a schematic illustrating the cascading dynamics governing the interaction between two input optical spectra separated by a frequency much larger than the phase-matched terahertz frequency is depicted.

 figure: Fig. 2

Fig. 2 Schematic describing the spectral dynamics of the present system. (a) A pump (higher intensity) and seed (lower intensity) with very disparate center frequencies. The frequency separation between pump and seed is very different fromthe phase-matched terahertz frequency. The angular frequency of the pump ωp may be larger or smaller than ωs. However, a few lines separated by the phase-matched terahertz angular frequency ΩTHz are centered about the seed at ωs, which beat to generate terahertz radiation. This subsequently mediates the cascaded processes. (b) The generated terahertz radiation first modulates both the pump and seed lines, producing a series of side-bands ωs±m and ωppmm (m = 1, 2, …), about the seed and pump respectively by DFG and SFG processes. (c) Eventually, the spectral gap between the pump and seed reduces and the process produces a net red-shift if phase matching conditions are favorable. This is accompanied with an increase in terahertz generation.

Download Full Size | PDF

The pump spectrum is narrowband, centered at angular frequency ωp and is represented by a single line in Fig. 2(a). Experimentally, this spectral distribution may correspond to a long pulse with a transform-limited duration of >100 ps. The seed is centered at angular frequency ωs and contains a set of lines separated by the phase-matched terahertz angular frequency ΩTHz. A minimum of two lines shall be required for the seed. Each seed line may correspond to a narrowband pulse, also of transform-limited duration >100 ps. Variants of this input pulse format are described at the end of this section and depicted in Fig. 3. The physical processes at work for those formats can be readily deduced based on the discussion in this paper.

Here and throughout this paper, the pump is defined to be the pulse (or line) with larger intensity. The angular frequency of the pump ωp may be at a smaller or larger frequency compared to the seed ωs. In terahertz generation systems, one of the peculiar consequences of cascading effects is that the initial conditions tend to rapidly wash out due to the simultaneous evolution of many strongly coupled three-wave mixing processes. This renders the relative location of pump and seed unimportant.

In the initial step as shown in Fig. 2(b), DFG between spectral components centered around ωs

produces some terahertz radiation at ΩTHz. Subsequently, the generated terahertz radiation interacts with the pump at ωp and ωs to produce side-bands ωp±1,ωp±2 and ωs±1,ωs±2 about both frequencies respectively. As previously mentioned, this is because both SFG and DFG processes are well phase matched due to the small value of ΩTHz compared to ωp or ωs.

If phase matching conditions for frequencies red-shifted with respect to the pump, i.e. for ω<ωp are more favorable, a preferential red-shift of the total (or combined) optical spectrum eventually occurs as shown in the final schematic in Fig. 2(c). This red-shift is accompanied with significant terahertz generation.

For the envisaged dynamics in Figs. 2(a)-(c), terahertz absorption would have to be low enough to enable sufficient interaction. In practice, this can be realized in cryogenically-cooled lithium niobate.

In subsequent sections, we will numerically prove that the spectral evolution follows the above description and furnish further details on the process.

 figure: Fig. 3

Fig. 3 Variations in input spectral formats producing similar dynamics. (a) Seed is located at a smaller angular frequency, i.e. ωs < ωp. (b) Additional lines to generate terahertz radiation are distributed about the pump rather than the seed, i.e. ωp±1 = ωp ± ΩTHz and ωp < ωs. (c) Additional lines are provided adjacent to the pump and ωp > ωs. In all cases, the cascading dynamics will be similar if phase-matching towards red-shifted frequencies is superior.

Download Full Size | PDF

In addition to the input spectral format described in Fig. 2(a), other variations depicted in Fig. 3 can also be utilized. In Fig. 3(a), the seed is situated at an angular frequency smaller than the pump, i.e. ωs<ωp. In Fig. 3(b), the additional lines to generate terahertz radiation are distributed about ωp rather than ωs. Finally, in Fig. 3(c), ωp<ωs and contains additional lines distributed about ωp. The cascading dynamics in all these cases are similar to Fig. 2 and require phase-matching to be superior in the direction of red-shift for efficient terahertz generation.

Tables Icon

Table 1. Parameters used in simulations

3. Results

3.1. Numerical model

To study the general physics governing the approach, a model which approximates each spectral line as being monochromatic is used. In reality, each line has a narrow bandwidth corresponding to a transform-limited duration of >100 ps. However, tracking the evolution of the continuous spectrum is computationally expensive. Therefore, we may employ the computationally efficient discrete approximation to furnish an overarching understanding of the physics governing the system. Furthermore, the discrete and continuous approaches are in qualitative as well as reasonable quantitative agreement, which justifies the discrete approximation [15, 16]. In the final section, we model the complete continuous spectral distribution without any approximations as in [27] for reliable quantitative predictions.

The terahertz spectral envelope as a function of longitudinal distance z is represented by ATHz(z), while the optical/NIR ones are represented by Am(z). Their evolution is presented in Eqs. (1a) and (1b) respectively.

dATHz(z)dz=α2ATHz(z)jΩ2χ(2)(z)2k(Ω)c2mAm+1(z)Am(z)*ej[km+1kmk(Ω)]z
dAm(z)dz=jωm2χ(2)(z)2kmc2[Am+1(z)ATHz*(z)ej[km+1kmk(Ω)]z+Am1(z)ATHz(z)ej[km1+k(Ω)km]z]
η(z)=nTHz|ATHz(Ω,z)|2Σmn(ωm)|Am(0)|2=nTHz|ATHz(Ω,z)|2Fpump+Fseed

In Eq. (1a), the first term represents terahertz absorption while the second represents the ensemble of DFG processes between all spectral components in the optical domain. The second-order nonlinear susceptibility is denoted by χ(2)(z), with the z dependence provided to account for poled electro-optic crystals. In Eq. (1b), the first term within square brackets represents beating of the m+1th spectral component in the optical domain and terahertz radiation to generate the spectral component m. The second term however represents SFG of the mth spectral component in the optical domain by addition of the frequencies of the m1th spectral component and terahertz radiation. The optical-to-terahertz conversion efficiency η(z) is given by Eq. (1c). The efficiency is calculated by taking the ratio of the generated terahertz fluence and the total optical fluence, i.e. of the pump and seed taken together.

To provide a quantitative illustration of the picture outlined in the previous section, we consider seed lines with an intensity distribution Is,m given by Eq. (2).

Is,m=Isexp[(ωm2πfc)2/(NwfTHz)2]

In Eq. (2), fc is the spectral separation between the two input spectra (pump and seed) while Nw represents the bandwidth of the seed distribution in integer units (or number of lines). Is denotes the peak intensity among all seed lines. In the current simulation, fc = 10 THz and Nw = 2. The terahertz frequency fTHz=ΩTHz/(2π) is set to be 0.5 THz. A single pump line is considered with Is=0.01Ip.

A PPLN crystal with 5%

Magnesium Oxide (MgO) doping is assumed to be phase matched for the generation of fTHz=0.5 THz under conditions of cryogenic cooling at T=100 K. Full material dispersion in the optical domain is obtained from [22], while that for the terahertz region is obtained from [24]. Terahertz absorption coefficients at cryogenic temperatures are obtained from [23]. Using these material properties, the required poling period is obtained as Λ=240μm. The second-order nonlinear susceptibility of bulk lithium niobate is assumed to be χ0(2)=336 pm/V. The various material parameters used in this paper are tabulated in Table 1.

Each line is assumed to correspond to a transform-limited e2 duration of τ = 200 ps. The total fluence, i.e. for all pump and seed lines put together is set to 0.7 Jcm 2. This corresponds to half the laser-induced damage threshold value, which is given by the relation Fd=1(τ/100ps)1/2J/cm2 [27].

 figure: Fig. 4

Fig. 4 (a) The total optical spectrum (i.e. pump and seed) is plotted for various locations along the crystal z in panels (i)-(iv). Initially, the pump and seed spectra are very disparate in panel (i). After initial modulation bythe generated terahertz radiation, they begin to merge as seen at location (ii). Subsequently, they completely merge in (iii), followed by significant preferential red-shift in panel (iv). The merging of the spectra, i.e. at z=4 cm is seen to be the point at which conversion efficiency begins to grow rapidly. (b) Terahertz efficiency growth as a function of propagation length z for fc = 10 THz, Is = 0.01Ip and Nw = 2 in cryogenically-cooled PPLN crystals. An exponential growth of terahertz conversion efficiency, followed by a drop due to phase mismatch is observed. (c) The phase-mismatch as a function of detuning from the signal frequency is plotted. Phase-matching for red shifting with respect to ωp is preferred.

Download Full Size | PDF

In Fig. 4(a), the total optical spectrum, i.e. of both pump and seed are depicted. The panels labeled (i)-(iv) delineate the total spectrum at different locations along the propagation direction z inside the crystal.

As can be seen, the initial input format consists of two spectra centered at angular frequencies ωs and ωp. The spectral distribution about ωs is of significantly lower intensity and contains lines separated by the terahertz frequency fTHz. Initially, the generated terahertz radiation is weak and functions primarily as a means to modulate the two input spectra, resulting in a gradual merging of the pump and seed spectra as can be seen in panel (ii) of Fig. 4(a). Only after z = 4 cm, does the modulation give way to a preferential red-shift as can be seen in panels (iii)-(iv) of Fig. 4(a).

In Fig. 4(b), we plot the terahertz conversion efficiency as a function of length. The arrows labeled (i)-(iv) correspond to the locations represented by panels (i)-(iv) in Fig. 4(a). No terahertz radiation is incident on the crystal and therefore the conversion efficiency η at z = 0 is zero. The conversion efficiency remains low at z = 2 cm, since the generated terahertz radiation has been primarily modulating the pump and seed spectra as evident in panel (ii) of Fig. 4(a). It is only at z4 cm that η begins to experience significant growth. This coincides with a preferential red-shift in the total spectrum as is seen in panel (iii) of Fig. 4(a). The terahertz conversion efficiency then grows exponentially over the crystal length, followed by a drop beyond z7 cm. This drop is due to phase mismatch which occurs as the total optical bandwidth increases by spectral broadening induced by cascading processes. In the presence of phase mismatch, terahertz absorption results in a drop in conversion efficiency.

Figure 4(c) depicts phase matching as a function of detuning from the pump frequency. As can be seen, phase mismatch is larger for the region of blue shift (Δf>0) compared to the region of red shift (Δf<0). This shall be shown to be a key requirement for growth in terahertz efficiency.

3.2. Effect of dispersion

To further understand the cascading process in this system, we examine its behavior under various conditions of dispersion. In this section, absorption is switched off to not obfuscate the overarching physics governing the process. Including absorption reduces the maximum efficiency that is reached and also results in a reduction in efficiency once there is significant phase mismatch as was evident in Fig. 4(b) beyond z7 cm. We revert to including absorption in subsequent sections.

 figure: Fig. 5

Fig. 5 Dispersion curves and their influence on cascading dynamics. Absorption is switched off to not obfuscate the overarching physics without loss of generality. (a) Plot showing evolution of the total optical spectrum along crystal length for conventional phase-matching conditions in PPLN. This represents the experimentally relevant case. Initial modulation of pump and seed spectra around ωp, ωs respectively is evident. Effective red-shift occurs only when the two spectra merge at ≈4 cm. This is followed by preferential red-shift and then subsequently back-conversion due to phase-mismatch. (b) Cascading dynamics for a fictitious dispersion curve where ωp < ω < ωs and ω > ωs are very highly phase mismatched. In this situation, the initially generated terahertz radiation by beating between lines distributed about ωs, drives the continuous red-shift of the spectrum about ωp. Since the need for modulation is greatly reduced, significant red-shift occurs at much shorter distances of z = 2 cm. (c) A dispersion curve, where only ω > ωs is phase-mismatched. In this case, the spectrum about ωp will blue shift to a much larger extent till it reaches ωs before commencing a red shift. Thus, the threshold point occurs at distances z = 6 cm, i.e. larger than that in (a). (d) Conversion efficiencies for various cases (a)-(c). (e) Dispersion in phase matching corresponding to the cases in (a)-(c).

Download Full Size | PDF

In Fig. 5(a), the total optical spectrum as a function of propagation distance z is plotted for the case when dispersion in phase matching is identical to that depicted in the previous section (blue curve in Fig. 5(e)). This represents the experimentally relevant case. In Fig. 5(b), the spectral dynamics for a situation where Δk=0 except for ωp<ω<ωs and for ω>ωs (red curve in Fig. 5(e)) is shown. Finally, in Fig. 5(c),Δk0 only for ω>ωs (black, Fig. 5(e)).

In Fig. 5(a), the terahertz radiation generated by DFG between lines distributed around the seed centered at ωs initially modulates both pump and seed spectra until preferential phase matching in the region red-shifted with respect to ømegas, i.e. Δf<0, begins to take precedence over blue-shifting. This occurs at z4 cm, which is when optical-to-terahertz conversion efficiency begins to grow as is evident in the blue curve in Fig. 5(d).

In Fig. 5(b), the region ωp<ω<ωs is so highly phase mismatched that effective red shifting occurs right from the outset, leading to much earlier growth in terahertz efficiency at z2 cm (red curve, Fig. 5(d)). This is because, blue-shifting with respect to the seed is prohibited by the phase mismatch in that region.

In Fig. 5(c), phase matching is violated only for ω>ωs. Therefore, the pump continues to be modulated until it spreads from ωp till it reaches the spectral distribution around ωs. At this point, when blue shifting can no longer occur, the red-shifted processes begin to dominate and terahertz efficiency growth occurs. Since the spectrum has to spread to a much greater extent compared to cases (a) and (b), the growth of terahertz conversion efficiency occurs much later at z6 cm as can be seen in Fig. 5(d) (black curve).

Thus from previous numerical simulations, it can be seen that the main requirement for a net red-shift is that phase matching must improve for red-shifted optical/NIR frequencies. A caveat however is that the distance along the propagation direction z at which preferential red shift occurs must be much larger than the terahertz absorption length. Otherwise, the terahertz field would be too strongly attenuated before the modulated spectrum reaches a point of large phase mismatch (in the blue-shifted region) to result in preferential red shift.

3.3. Effect of Nw

In this section, we investigate the effect of the bandwidth of the seed Nw (or loosely the number of seed lines) on the cascading dynamics in Fig. 6. The role of the frequency separation fc between the pump and seed spectra can be similarly deduced. In the ensuing simulations, the exact material parameters for lithium niobate at T=100 K from Table 1 are assumed. We revert to including the effects of absorption in this section and all subsequent sections. A PPLN with Λ=240μm, corresponding to a phase-matched terahertz frequency of 0.5 THz is considered.

In general, increasing Nw should result in the faster onset of preferential red shift. This is because the distance at which the two input spectra are modulated to an extent that they begin to overlap and force a net red shift decreases. This can be seen by comparing Fig. 6(a) to Fig. 5(a). In Fig. 6(a), fc = 10 THz as in Fig. 5(a) but Nw is increased to 5 as opposed to Nw = 2 in Fig. 5(a).

This is further evident upon examining the terahertz conversion efficiency as a function of propagation distance z in Fig. 6(b). Here, the cases from Fig. 6(a) (Nw=5,fc=10 THz) experiences exponential growth at a shorter distance of z = 3 cm compared to the case when Nw=2,fc=10 THz (black-dotted curve in Fig. 6(b)).

Decreasing fc would have a similar effect to increasing Nw since it too would result in an effective reduction of the gap between the pump and seed spectra.

 figure: Fig. 6

Fig. 6 (a) Spectral dynamics for Nw = 5 lines about ωs separated by fc = 10 THz from ωp. (b) The threshold point at which terahertz efficiency experiences growth is reduced compared to the case of Nw = 2, fc = 10 THz.

Download Full Size | PDF

From the above set of numerical experiments in Figs. 46, we have deduced that phase matching for frequencies lower than the pump must be better in the case when ωp<ωs. Furthermore, we find that the onset of efficiency growth occurs only when the pump and seed spectra begin to overlap.

3.4. Effect of terahertz frequency and seed intensity

The above understanding might be extended to other relevant parameters such as the pump intensity or terahertz frequency. We may anticipate that larger terahertz frequencies close the gap faster and would hence experience terahertz growth at shorter distances. Furthermore, since conversion efficiency increases with terahertz frequency, larger frequencies may be anticipated to be generated with larger conversion efficiencies. However, the increase is not incessant due to the effects of absorption which also increase with frequency. This is evident in Fig. 7(a), which depicts the maximum conversion efficiency for various phase-matched terahertz frequencies. The material properties are once again obtained from Table 1. The PPLN period in each case is defined by Λ=c(nTHz(fTHz)ng)1/fTHz, where nTHz is the terahertz phase index and ng is the optical group refractive index. In these simulations, fc = 10 THz while Nw = 2. The peak intensity of the seed is set to be Is=0.01Ip and the total fluence to 0.7Jcm2.

An increase in conversion efficiency with terahertz frequency followed by saturation at 0.5 THz is observed in Fig. 7(a). Furthermore as anticipated, the optimal lengths reduce for larger phase-matched terahertz frequencies. A larger seed or pump intensity would expedite the process and also result in shorter interaction lengths (red curve Fig. 7(b)).

 figure: Fig. 7

Fig. 7 (a) Terahertz conversion efficiency as function of terahertz frequency shows initial increase and subsequent saturation due to increased absorption. (b) Requisite crystal lengths for maximizing efficiency show inverse dependence with terahertz frequency and peak seed intensity.

Download Full Size | PDF

3.5. Temporal evolution and damage thresholds

In this section, we examine the temporal evolution of the total optical field, i.e. of both pump and seed put together. This provides guidelines for optimizing the input spectrum to mitigate laser-induced damage.

In Fig. 8, we plot the spectral and temporal evolution of various input spectra along the length of the crystal. The spectral and temporal evolution up until the point when maximum optical intensity is reached along the crystal length is depicted. The temporal envelope of the total optical spectrum is given by E(t,z)=Re{ΣmAm(z)ejkmzet2/τ2}, where Am corresponds to the mth optical line whose evolution is delineated in Eq. (1b). The panelson the left of Fig. 8 represent the total optical spectrum at various values of z, while the right hand panels represent the corresponding intensity profiles I(t) in the time-domain.

In Fig. 8(a), we plot the spectral and temporal evolution for the pulse format utilized in Fig. 4(a) and 5(a). It is comprised of a pump separated by fc = 10 THz from the seed with bandwidth Nw = 2 in Eq. (2). The peak seed intensity is given by Is=0.01Ip. The total fluence of pump and seed taken together is set to half the laser-induced damage threshold value, i.e. Fd/2=0.7Jcm2. A PPLN crystal phase-matched for the terahertz frequency fTHz=0.5 THz at T=100 K is assumed. Material and other properties are obtained from Table 1.

As evident from the first panel on the right hand side of Fig. 8(a), at z = 0, the combination of pump and seed represent an envelope with e2 duration τ = 200 ps. The envelope is modulated at a rate of 0.5 THz or with a temporal period of 2 ps. The total optical spectral bandwidth increases along the crystal length due to cascading effects as evident in the left hand side panels of Fig. 8(a). In the time-domain, the modulations get deeper and eventually a pulse train with a period of 2 ps is obtained (see inset of final panel on the right hand side of Fig. 8(a)). Importantly, one notices a dramatic increase in peak intensity to >25GWcm2 compared to the initial peak intensity of 5GWcm2 which is an issue from the point of view of laser-induced damage.

Since the intensity spiking behavior arises from a large increase in spectral bandwidth, it is presumable that starting with a larger number of lines around the seed centered at angular frequency ωs would result in reduced intensity spiking. This is seen in Fig. 8(b), where Nw = 5 and Is=0.01Ip. As in Fig. 8(a), the total fluence is maintained at half the laser-induced damage threshold value of Fd/2=0.7Jcm2. It is evident in contrasting the first and last panels on the right-hand side of Fig. 8(b) that the intensity increase is only threefold as opposed to being more than fivefold in Fig. 8(a). However, the peak intensity at z=5.5 cm in the last panel on the right hand side of Fig. 8(b) is 30GWcm2 which is larger than the value of 25GWcm2 in Fig. 8(a). This can be remedied as shown in Fig. 8(c).

In Fig. 8(c), we increase the fraction of the seed to pump intensity, i.e. Is/Ip to being 10% rather than 1%. However, the total input fluence has to be reduced to maintain the peak intensity below reasonable levels. This is achieved by setting the total fluence to Fd/4=0.35Jcm2, which is half that in Figs. 8(a)8(b). We find that in this case, even as cascading occurs, the intensity level is roughly maintained at 10GWcm2 in the right hand panels of Fig. 8(c). To ascertain exactly what peak intensity is permissible, we track free-electron densities in the subsequent section.

 figure: Fig. 8

Fig. 8 Spectral (left hand side panels) and temporal evolution (right hand side panels) along crystal length z for various input pulse formats in (a)-(c). A larger initial bandwidth of seed (i.e. larger Nw) and larger ratio of Is/Ip enables keeping the peak intensity stable during the cascading process. However, while intensity growth is reduced in such cases, absolute intensity may still be higher as seen in (b). To circumvent this, the total input fluence can be reduced as in (c). The maximum permissible intensity within the crystal is ascertained by tracking the generated free-electron density in the subsequent section.

Download Full Size | PDF

3.6. Free-electron density

To determine what might be considered a reasonable input fluence, we use the generated free-electron density in the material as an indicator of the onset of laser-induced damage. This approach assumes the widely used avalanche-breakdown model for laser-induced damage [28]. The free-electron density Nc(t,z) is obtained according to Eq. (3).

dNc(t,z)dt=gavI(t,z)Nc(t)+Eg1β(4)I4(t,z)

In Eq. (3), the first term represents the generation of free electrons by an avalanche process. Here, gav is the avalanche coefficient while I(t) is the temporal intensity profile corresponding to the combined pump and seed spectra. The second term in Eq. (3) corresponds to the generation of free electrons by four-photon absorption of photons in the 1μm wavelength region. Here, Eg is the bandgap of the material, while β(4) is the four-photon absorption coefficient.

For lithium niobate and input radiation in the 1μm region, a minimum of four photons would need to be absorbed since the bandgap of lithium niobate is Eg4 eV. However, for input radiation centered around 800 nm, even three-photon absorption would have to be considered. Higher-order absorption coefficients are neglected since these events are of much lower probability for the typical intensity levels characterizing long pulses considered in the current system.

It is worth noting that the primary impact of multi-photon absorption in the current system is as a source of free-electrons which may then initiate a breakdown process via avalanche effects. However, the impact of multi-photon absorption in attenuating the pump and/or seed can expected to be minimal. This is because the absorption coefficient β(4)I3 due to four-photon absorption is only 1.7×102m1 for a peak intensity of 10 GWcm 2. This would translate to a transmission of 99.9% over a 6 cm long crystal. Thus, multi-photon absorption is not expected to attenuate the pump or seed appreciably.

Using parameters from Table 1, we plot the optical-to-terahertz conversion efficiency growth as a function of crystal length for various pulse formats in Fig. 9 (a) along with the maximum free-electron density (over time) along crystal length in 9(b).

As can be seen, for Nw = 2, Is/Ip=0.01 and a total input fluence of Fd/2, there is an exponential growth in the conversion efficiency η beyond z4 cm. The increase in conversion efficiency is known to be accompanied by an increase in bandwidth due to cascading from Fig. 8(a). This in turn results in a spiking of optical intensity thereby resulting in an exponential growth in free-electron density in Fig. 9(b) (black-dotted curve). The free-electron density rises above the threshold value (delineated by the green horizontal line) significantly. The threshold free-electron density is evaluated using a single pulse with e2 pulse duration τ = 200 ps and fluence given by the laser-induced damage threshold value from Table 1.

Increasing the ratio of seed to pump intensity Is/Ip to 10% and Nw to 5 alleviates temporal intensity growth. This in turn, eliminates the drastic increase in free-electron density along the crystal length characterizing the Nw=2,Is=0.01Ipcase. This is seen in the red curve of Fig. 9(b). However, the value of the free-electron density is larger than the threshold (green) even at the outset.

 figure: Fig. 9

Fig. 9 (a)Terahertz efficiency versus length for various values of Is/Ip, number of seed lines Nw and total fluence levels. (b) Corresponding carrier density generated for each case as a function of length. The green horizontal line represents the laser-induced damage threshold free-electron density calculated by using the laser-induced damage threshold fluence levels from Table 1.

Download Full Size | PDF

If we reduce the total input fluence for the Nw=5,Is=0.1Ip case by a factor of 2 to Fd/4, we retain the stability of intensity but also keep the free-electron density below the threshold value throughout (blue curve in Fig. 9(b)).

At the same time, the conversion efficiency reduces from approximately 4.5% (red-solid, Fig. 9(a)) to 3% (blue, Fig. 9(a)). Therefore, one can still obtain conversion efficiencies between 34.5% while addressing the intensity spiking behavior due to increased bandwidth of the total spectrum.

Thus, in this section we’ve established that having a larger initial seed bandwidth can solve the intensity spiking behavior but will possess a lower effective damage fluence by virtue of larger peak intensities at the outset. This can be mitigated by reducing the total input fluence, while still obtaining high conversion efficiency, albeit lower.

4. Application: Terahertz generation

In this section, we evaluate terahertz generation with the pump located at λp=1064 nm and seed lines centered about λs=1030 nm. A cryogenically-cooled PPLN crystal at T=80 K is assumed. The phase-matched terahertz frequency is set to 0.3 THz due to its relevance to compact electron accelerator technology [21]. The choice of λp,λs is informed by their suitability to being amplified by high energy Nd:YAG and Yb:YAG amplifiers respectively.

4.1. Experimental setup

In practice, such a source can be realized by first locking laser lines at λp=1064 nm and λs=1030 nm to a frequency comb whose repetition rate frepfTHz [29]. This allows for the programmable generation of phase-stable arbitrary spectra as was recently demonstrated [29]. These phase-locked spectra can then be amplified to very high energies in separate laser amplifiers. For instance, the phase-locked pump line at λp=1064 nm maybe amplified using Nd:YAG amplifiers while the phase-locked lines centered at λs=1030 nm can be amplified by high energy, high repetition rate Yb:YAG amplifiers [30].

4.2. Results

The pump and seed lines are all assumed to correspond to pulses of τ = 200 ps e2 duration without loss of generality. For longer durations of ∼ns, a reduced conversion efficiency would be obtained which can be mitigated by recycling the cascaded optical spectrum in multiple crystals [31].

We perform simulations without the discrete approximations of Eqs. (1a)(1b) based on the model presented in [27] and parameters from Table 1. The key equations are as follows:

dATHz(Ω,z)dz=α(Ω)2ATHz(Ω,z)jΩ2χ(2)(z)2k(Ω)c20Aop(ω+Ω,z)Aop*(ω,z)ej[k(ω+Ω)k(ω)k(Ω)]zdω
dAop(ω,z)dz=jω2χ(2)(z)2k(ω)c2[0Aop(ω+Ω,z)ATHz*(Ω,z)ej[k(ω+Ω)k(ω)k(Ω)]zdΩ+0Aop(ωΩ,z)ATHz(Ω,z)ej[k(ωΩ+k(Ω)k(ω)]zdΩ]jε0ω0n(ω0)n22F[|Aop(t)|2Aop(t)]
η(z)=πcε00nTHz(Ω)|ATHz(Ω,z)|2dΩFpump+Fseed

Equation (4a) is analogous to Eq. (1a) but considers the continuous terahertz spectrum ATHz(Ω,z). Notice that the second term in Eq. (4a) represents the ensemble of all DFG processes and is an integral representation of the second term in Eq. (1a). Similarly, Eq. (4a) is analogous to Eq. (1b). Here, Aop(ω,z) is used to encompass the complete optical spectrum. The first two terms of Eq. (4a) are analogous to the first two terms of Eq. (1b). The third term in Eq. (4b) represents self-phase modulation with nonlinear refractive index n2. For a more detailed explanation of Eqs. (4a)(4b), the reader is referred to [27]. Equation (4c) is analogous to Eq. (1c) for the calculation of the optical-to-terahertz conversion efficiency η(z).

 figure: Fig. 10

Fig. 10 (a)Terahertz efficiency versus length in PPLN crystals phase matched for 0.3 THz for various values of Is/Ip,Nw. The pump is located at 1064 nm and seed at 1030 nm. (b) Spectral broadening and red-shift accompanying the percent-level conversion efficiencies.

Download Full Size | PDF

In Fig. 10(a), the conversion efficiency for various values of Nw and Is/Ip are presented. The total fluence is set to be 0.7Jcm2, which is a factor of two lower than the laser-induced damage threshold fluence Fd specifiedin Table 1. For cm 2-aperture PPLN crystals [17], this would permit pump energies of >0.5 J. The optimization of fluence and input spectrum to mitigate laser-induced damage is not undertaken here for the sake of brevity. However, it would follow along the lines of section 3.6 and shall be reported elsewhere.

As described in section 3.4, the cascading process occurs more gradually with distance for lower terahertz frequencies. Therefore, in relation to the 0.5 THz cases previously discussed, conversion efficiency growth for the 0.3 THz is slower with respect to crystal length. In Fig. 10(a), for Is/Ip=0.01 the conversion efficiency is quite low for various Nw even after 6 cm of propagation. In principle, one may use multiple PPLN crystals and recycle the cascaded optical spectrum to circumvent this issue [31].

However, if high efficiencies at the percent level need to be obtained at shorter crystal lengths, the ratio has to be increased to Is=0.1Ip as seen in Fig. 10(b). In Fig. 10(a), for Nw = 10 (black curve), the drop in efficiency is attributed to phase mismatch after the optical spectrum broadens significantly.

In Fig. 10(b), the optical spectrum at various z locations for the case when Nw=10,Is=0.1Ip are presented. An increasing amount of red shift after merging of the pump and seed spectra, consistent with the previously developed physical picture is seen.

For the percent-level conversion efficiencies potentially achievable using these pulse formats and Joule-level pump energies, narrowband terahertz pulses with several millijoules of pulse energy thus appear feasible.

5. Conclusion

The spectral dynamics in the case of a system with pump and seed spectra separated by a large frequency offset is discussed when one of the pump or seed contains bandwidths larger than the terahertz frequency to be generated. It is important to note that the frequency offset can be very different from the phase-matched terahertz frequency, which is contrary to our previous work [15, 16]. In such a case, the phase-matched terahertz wave thus generated mediates the cascaded interaction between the pump and seed spectra. This can eventually result in the merging of the pump and seed spectra. For conditions of reducing phase-mismatch in the direction of the desired red-shift, high optical-to-terahertz conversion efficiency can be achieved. The issues of intensity spiking that accompany cascading effects was highlighted. Mitigating mechanisms using more broadband seed spectra was suggested. The approach was shown to be very valuable to generating narrowband terahertz radiation at percent-level conversion efficiencies and millijoule-level pulse energies using available 1064 nm high energy lasers in combination with 1030 nm seed pulses. The approach could be instrumental for enabling terahertz energy hungry applications such as electron acceleration and X-ray generation.

Funding

Air Force Office of Scientific Research (AFOSR-A9550-12-1-0499); H2020 European Research Council (609920); Seventh Framework Programme (FP/2007-2013).

References

1. R. Schiek, Y. Baek, and G. I. Stegeman, “One-dimensional spatial solitary waves due to cascaded second-order nonlinearities in planar waveguides,” Phys. Rev. E 53, 1138 (1996). [CrossRef]  

2. A. Varanavičius, A. Dubietis, A. Beržanskis, R. Danielius, and A. Piskarskas, “Near-degenerate cascaded four-wave mixing in an optical parametric amplifier,” Opt. letters 22, 1603–1605 (1997). [CrossRef]  

3. M. Cronin-Golomb, “Cascaded nonlinear difference-frequency generation of enhanced terahertz wave production,” Opt. Lett. 29, 2046–2048 (2004). [CrossRef]   [PubMed]  

4. J. Moses and F. W. Wise, “Soliton compression in quadratic media: high-energy few-cycle pulses with a frequency-doubling crystal,” Opt. Lett. 31, 1881–1883 (2006). [CrossRef]   [PubMed]  

5. M. Bache, J. Moses, and F. W. Wise, “Scaling laws for soliton pulse compression by cascaded quadratic nonlinearities,” JOSA B 24, 2752–2762 (2007). [CrossRef]  

6. M. Bache, “Cascaded nonlinearities for ultrafast nonlinear optical science and applications,” Ph.D. thesis, Danmarks Tekniske Universitet (2017).

7. K. L. Vodopyanov, “Optical generation of narrow-band terahertz packets in periodically-inverted electro-optic crystals: conversion efficiency and optimal laser pulse format,” Opt. Express 14, 2263–2276 (2006). [CrossRef]   [PubMed]  

8. K. Vodopyanov, “Optical thz-wave generation with periodically-inverted gaas,” Laser & Photonics Rev. 2, 11–25 (2008). [CrossRef]  

9. M. Jewariya, M. Nagai, and K. Tanaka, “Enhancement of terahertz wave generation by cascaded χ(2) processes in linbo3,” J. Opt. Soc. Am. B 26, A101–A106 (2009). [CrossRef]  

10. D. A. Walsh, P. G. Browne, M. H. Dunn, and C. F. Rae, “Intracavity parametric generation of nanosecond terahertz radiation using quasi-phase-matching,” Opt. express 18, 13951–13963 (2010). [CrossRef]   [PubMed]  

11. D. Molter, M. Theuer, and R. Beigang, “Nanosecond terahertz optical parametric oscillator with a novel quasi phase matching scheme in lithium niobate,” Opt. Express 17, 6623–6628 (2009). [CrossRef]   [PubMed]  

12. S. R. Tripathi, Y. Taira, S. Hayashi, K. Nawata, K. Murate, H. Minamide, and K. Kawase, “Terahertz wave parametric amplifier,” Opt. Lett. 39, 1649–1652 (2014). [CrossRef]   [PubMed]  

13. K. Murate and K. Kawase, “Perspective: Terahertz wave parametric generator and its applications,” J. Appl. Phys. 124, 160901 (2018). [CrossRef]  

14. G. Cirmi, M. Hemmer, K. Ravi, F. Reichert, L. E. Zapata, A.-L. Calendron, H. Çankaya, F. Ahr, O. D. Mücke, N. H. Matlis, and F. X. Kärtner, “Cascaded second-order processes for the efficient generation of narrowband terahertz radiation,” J. Phys. B: At. Mol. Opt. Phys. 50, 044002 (2017). [CrossRef]  

15. M. Hemmer, G. Cirmi, K. Ravi, F. Reichert, F. Ahr, L. Zapata, O. D. Mücke, A. Calendron, H. Çankaya, D. Schimpf, N. H. Matlis, and F. X. Kärtner, “Cascaded interactions mediated by terahertz radiation,” Opt. Express 26, 12536–12546 (2018). [CrossRef]   [PubMed]  

16. K. Ravi, M. Hemmer, G. Cirmi, F. Reichert, D. N. Schimpf, O. D. Mücke, and F. X. Kärtner, “Cascaded parametric amplification for highly efficient terahertz generation,” Opt. letters 41, 3806–3809 (2016). [CrossRef]  

17. H. Ishizuki and T. Taira, “Improvement of laser-beam distortion in large-aperture ppmgln device by using x-axis czochralski-grown crystal,” Opt. Express 22, 19668–19673 (2014). [CrossRef]   [PubMed]  

18. E. A. Nanni, W. R. Huang, K.-H. Hong, K. Ravi, A. Fallahi, G. Moriena, R. J. Dwayne Miller, and F. X. Kärtner, “Terahertz-driven linear electron acceleration,” Nat. Commun. 6, 8486 (2015). [CrossRef]   [PubMed]  

19. D. Zhang, A. Fallahi, M. Hemmer, X. Wu, M. Fakhari, Y. Hua, H. Cankaya, A.-L. Calendron, L. E. Zapata, N. H. Matlis, and F. X. Kärtner, “Segmented terahertz electron accelerator and manipulator (steam),” Nat. Photonics 12, 336–442 (2018). [CrossRef]   [PubMed]  

20. F. Lemery, K. Floettmann, P. Piot, F. Kärtner, and R. Aßmann, “Synchronous acceleration with tapered dielectric-lined waveguides,” Phys. Rev. Accel. Beams 21, 051302 (2018). [CrossRef]  

21. F. Kärtner, F. Ahr, A.-L. Calendron, H. Çankaya, S. Carbajo, G. Chang, G. Cirmi, K. Dörner, U. Dorda, A. Fallahi, A. Hartin, M. Hemmer, R. Hobbs, Y. Hua, W. Huang, R. Letrun, N. Matlis, V. Mazalova, O. Mücke, E. Nanni, W. Putnam, J. Ravi, F. Reichert, I. Sarrou, X. Wu, A. Yahaghi, Y. Hong, L. Zapata, D. Zhang, C. Zhou, R. Miller, K. Berggren, H. Graafsma, A. Meents, R. Assmann, H. Chapman, and P. Fromme, “Axsis: Exploring the frontiers in attosecond x-ray science, imaging and spectroscopy,” Nucl. Instruments Methods Phys. Res. Sect. A: Accel. Spectrometers. Detect. Assoc. Equip. 829, 24–29 (2016). [CrossRef]  

22. D. H. Jundt, “Temperature-dependent sellmeier equation for the index of refraction, n e, in congruent lithium niobate,” Opt. Lett. 22, 1553–1555 (1997). [CrossRef]  

23. J. A. Fülöp, L. Pálfalvi, M. C. Hoffmann, and J. Hebling, “Towards generation of mj-level ultrashort thz pulses by optical rectification,” Opt. Express 19, 15090–15097 (2011). [CrossRef]   [PubMed]  

24. L. Pálfalvi, J. Hebling, J. Kuhl, A. Peter, and K. Polgár, “Temperature dependence of the absorption and refraction of mg-doped congruent and stoichiometric linbo3 in the thz range,” J. Appl. Phys. 97, 123505 (2005). [CrossRef]  

25. Q. Meng, Z. Su, J. Yu, and B. Zhang, “Effect of major factors on damage threshold of optical rectification crystals,” in Optomechanical Engineering 2015, vol. 9573 (International Society for Optics and Photonics, 2015), p. 957305.

26. M. C. Hoffmann, K.-L. Yeh, J. Hebling, and K. A. Nelson, “Efficient terahertz generation by optical rectification at 1035 nm,” Opt. Express 15, 11706–11713 (2007). [CrossRef]   [PubMed]  

27. K. Ravi, D. N. Schimpf, and F. X. Kärtner, “Pulse sequences for efficient multi-cycle terahertz generation in periodically poled lithium niobate,” Opt. express 24, 25582–25607 (2016). [CrossRef]   [PubMed]  

28. R. W. Boyd, Nonlinear optics(Elsevier, 2003).

29. D. N. Schimpf, H. T. Olgun, A. Kalaydzhyan, Y. Hua, N. H. Matlis, and F. X. Kärtner, “Frequency-comb-based laser system producing stable optical beat pulses with picosecond durations suitable for high-precision multi-cycle terahertz-wave generation and rapid detection,” Opt. Express 27, 11037–11056 (2019). [CrossRef]   [PubMed]  

30. L. E. Zapata, H. Lin, A.-L. Calendron, H. Cankaya, M. Hemmer, F. Reichert, W. R. Huang, E. Granados, K.-H. Hong, and F. X. Kärtner, “Cryogenic yb: Yag composite-thin-disk for high energy and average power amplifiers,” Opt. letters 40, 2610–2613 (2015). [CrossRef]  

31. L. Wang, A. Fallahi, K. Ravi, and F. Kärtner, “High efficiency terahertz generation in a multi-stage system,” Opt. express 26, 29744–29768 (2018). [CrossRef]   [PubMed]  

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 (10)

Fig. 1
Fig. 1 Example of a previously studied terahertz generation system where it is necessary to consider the simultaneous evolution of many phase-matched second-order processes even if optical-to-terahertz energy conversion efficiencies are low. (a) Initial interactions in a cascaded terahertz parametric amplifier. An OPA-like input of a strong optical pump at ω0 and weaker optical seed at ω−1 is used. The frequency separation between pump and seed is required to be equal to the phase-matched terahertz frequency, i.e. ω0ω1 = ΩTHz. The double arrows delineate the pair of waves involved in the interaction and the dotted arrows indicate the product of those interactions. For an input comprised of a strong pump at ω0 and weak seed at ω1 in the optical domain, a terahertz wave at ΩTHz corresponding to the beat frequency is first generated which then drives via sum frequency generation, the component at ω+1. (b) In subsequent steps, the generated terahertz radiation drives generation of red-shifted components ω2,3 and further blue-shifted components, e.g. ω+2. SFG consumes a terahertz photon while DFG generates one. Therefore, absolute conversion efficiencies are low in the absence of a mechanism which prefers DFG.
Fig. 2
Fig. 2 Schematic describing the spectral dynamics of the present system. (a) A pump (higher intensity) and seed (lower intensity) with very disparate center frequencies. The frequency separation between pump and seed is very different fromthe phase-matched terahertz frequency. The angular frequency of the pump ωp may be larger or smaller than ωs. However, a few lines separated by the phase-matched terahertz angular frequency ΩTHz are centered about the seed at ωs, which beat to generate terahertz radiation. This subsequently mediates the cascaded processes. (b) The generated terahertz radiation first modulates both the pump and seed lines, producing a series of side-bands ωs±m and ωppmm (m = 1, 2, …), about the seed and pump respectively by DFG and SFG processes. (c) Eventually, the spectral gap between the pump and seed reduces and the process produces a net red-shift if phase matching conditions are favorable. This is accompanied with an increase in terahertz generation.
Fig. 3
Fig. 3 Variations in input spectral formats producing similar dynamics. (a) Seed is located at a smaller angular frequency, i.e. ωs < ωp. (b) Additional lines to generate terahertz radiation are distributed about the pump rather than the seed, i.e. ωp±1 = ωp ± ΩTHz and ωp < ωs. (c) Additional lines are provided adjacent to the pump and ωp > ωs. In all cases, the cascading dynamics will be similar if phase-matching towards red-shifted frequencies is superior.
Fig. 4
Fig. 4 (a) The total optical spectrum (i.e. pump and seed) is plotted for various locations along the crystal z in panels (i)-(iv). Initially, the pump and seed spectra are very disparate in panel (i). After initial modulation bythe generated terahertz radiation, they begin to merge as seen at location (ii). Subsequently, they completely merge in (iii), followed by significant preferential red-shift in panel (iv). The merging of the spectra, i.e. at z=4 cm is seen to be the point at which conversion efficiency begins to grow rapidly. (b) Terahertz efficiency growth as a function of propagation length z for fc = 10 THz, Is = 0.01Ip and Nw = 2 in cryogenically-cooled PPLN crystals. An exponential growth of terahertz conversion efficiency, followed by a drop due to phase mismatch is observed. (c) The phase-mismatch as a function of detuning from the signal frequency is plotted. Phase-matching for red shifting with respect to ωp is preferred.
Fig. 5
Fig. 5 Dispersion curves and their influence on cascading dynamics. Absorption is switched off to not obfuscate the overarching physics without loss of generality. (a) Plot showing evolution of the total optical spectrum along crystal length for conventional phase-matching conditions in PPLN. This represents the experimentally relevant case. Initial modulation of pump and seed spectra around ωp, ωs respectively is evident. Effective red-shift occurs only when the two spectra merge at ≈4 cm. This is followed by preferential red-shift and then subsequently back-conversion due to phase-mismatch. (b) Cascading dynamics for a fictitious dispersion curve where ωp < ω < ωs and ω > ωs are very highly phase mismatched. In this situation, the initially generated terahertz radiation by beating between lines distributed about ωs, drives the continuous red-shift of the spectrum about ωp. Since the need for modulation is greatly reduced, significant red-shift occurs at much shorter distances of z = 2 cm. (c) A dispersion curve, where only ω > ωs is phase-mismatched. In this case, the spectrum about ωp will blue shift to a much larger extent till it reaches ωs before commencing a red shift. Thus, the threshold point occurs at distances z = 6 cm, i.e. larger than that in (a). (d) Conversion efficiencies for various cases (a)-(c). (e) Dispersion in phase matching corresponding to the cases in (a)-(c).
Fig. 6
Fig. 6 (a) Spectral dynamics for Nw = 5 lines about ωs separated by fc = 10 THz from ωp. (b) The threshold point at which terahertz efficiency experiences growth is reduced compared to the case of Nw = 2, fc = 10 THz.
Fig. 7
Fig. 7 (a) Terahertz conversion efficiency as function of terahertz frequency shows initial increase and subsequent saturation due to increased absorption. (b) Requisite crystal lengths for maximizing efficiency show inverse dependence with terahertz frequency and peak seed intensity.
Fig. 8
Fig. 8 Spectral (left hand side panels) and temporal evolution (right hand side panels) along crystal length z for various input pulse formats in (a)-(c). A larger initial bandwidth of seed (i.e. larger Nw) and larger ratio of Is/Ip enables keeping the peak intensity stable during the cascading process. However, while intensity growth is reduced in such cases, absolute intensity may still be higher as seen in (b). To circumvent this, the total input fluence can be reduced as in (c). The maximum permissible intensity within the crystal is ascertained by tracking the generated free-electron density in the subsequent section.
Fig. 9
Fig. 9 (a)Terahertz efficiency versus length for various values of Is/Ip, number of seed lines Nw and total fluence levels. (b) Corresponding carrier density generated for each case as a function of length. The green horizontal line represents the laser-induced damage threshold free-electron density calculated by using the laser-induced damage threshold fluence levels from Table 1.
Fig. 10
Fig. 10 (a)Terahertz efficiency versus length in PPLN crystals phase matched for 0.3 THz for various values of I s / I p , N w. The pump is located at 1064 nm and seed at 1030 nm. (b) Spectral broadening and red-shift accompanying the percent-level conversion efficiencies.

Tables (1)

Tables Icon

Table 1 Parameters used in simulations

Equations (8)

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

d A T H z ( z ) d z = α 2 A T H z ( z ) j Ω 2 χ ( 2 ) ( z ) 2 k ( Ω ) c 2 m A m + 1 ( z ) A m ( z ) * e j [ k m + 1 k m k ( Ω ) ] z
d A m ( z ) d z = j ω m 2 χ ( 2 ) ( z ) 2 k m c 2 [ A m + 1 ( z ) A T H z * ( z ) e j [ k m + 1 k m k ( Ω ) ] z + A m 1 ( z ) A T H z ( z ) e j [ k m 1 + k ( Ω ) k m ] z ]
η ( z ) = n T H z | A T H z ( Ω , z ) | 2 Σ m n ( ω m ) | A m ( 0 ) | 2 = n T H z | A T H z ( Ω , z ) | 2 F p u m p + F s e e d
I s , m = I s exp [ ( ω m 2 π f c ) 2 / ( N w f T H z ) 2 ]
d N c ( t , z ) d t = g a v I ( t , z ) N c ( t ) + E g 1 β ( 4 ) I 4 ( t , z )
d A T H z ( Ω , z ) d z = α ( Ω ) 2 A T H z ( Ω , z ) j Ω 2 χ ( 2 ) ( z ) 2 k ( Ω ) c 2 0 A o p ( ω + Ω , z ) A o p * ( ω , z ) e j [ k ( ω + Ω ) k ( ω ) k ( Ω ) ] z d ω
d A o p ( ω , z ) d z = j ω 2 χ ( 2 ) ( z ) 2 k ( ω ) c 2 [ 0 A o p ( ω + Ω , z ) A T H z * ( Ω , z ) e j [ k ( ω + Ω ) k ( ω ) k ( Ω ) ] z d Ω + 0 A o p ( ω Ω , z ) A T H z ( Ω , z ) e j [ k ( ω Ω + k ( Ω ) k ( ω ) ] z d Ω ] j ε 0 ω 0 n ( ω 0 ) n 2 2 F [ | A o p ( t ) | 2 A o p ( t ) ]
η ( z ) = π c ε 0 0 n T H z ( Ω ) | A T H z ( Ω , z ) | 2 d Ω F p u m p + F s e e d
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.