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

Mode conversion efficiency to Laguerre-Gaussian OAM modes using spiral phase optics

Open Access Open Access

Abstract

An analytical model for the conversion efficiency from a TEM00 mode to an arbitrary Laguerre-Gaussian (LG) mode with null radial index spiral phase optics is presented. We extend this model to include the effects of stepped spiral phase optics, spiral phase optics of non-integer topological charge, and the reduction in conversion efficiency due to broad laser bandwidth. We find that through optimization, an optimal beam waist ratio of the input and output modes exists and is dependent upon the output azimuthal mode number.

© 2017 Optical Society of America

1. Introduction

Over the past 25 years there has been growing interest in exploiting the orbital angular momentum (OAM) properties of light [1]. Studies into the generation and uses of laser generated OAM beams including cold atom trapping [2,3], communications [4,5], and enhanced imaging techniques [6,7] have provided a rich area of research. Recently, interest has grown with the application of OAM in laser-plasma interactions [8,9] corresponding to particle guiding [10,11] and magnetic field generation [12–15]. Additionally, applications of OAM modes in high harmonic generation have gained much interest in the last few years [16,17]. Laser beams carrying OAM can be described using Laguerre-Gaussian (LG) modes that propagate with a rotating Poynting vector about the beam axis. These LG modes are characterized by two numbers, the azimuthal and radial integers (denoted and p respectively). An infinite number of orthogonal modes can be generated and propagated in theory, but in practice the mode structure and purity is limited by the generation technique.

Since Allen’s paper [1], many methods for generating OAM modes have been introduced and developed; from mixing and converting various Hermite-Gaussian modes [18], to diffractive optical elements [19]. Spatial light modulators are a popular choice for high purity OAM mode generation due to their flexibility to generate any mode on demand, but are costly and only useful for lower power applications [4]. Generation of OAM modes by spiral phase plate (SPP), introduced by Beijersbergen in 1994 [20], has attracted a lot of interest due to its mode conversion quality, cost, and ease of use. The ability to convert a Gaussian TEM00 mode into a distribution of LG modes strongly centered around a single mode makes the SPP extremely attractive to users of high powered lasers who cannot easily modify the cavity modes of the laser front end or intermediate laser amplifier stages. Such phase plates can be designed to meet any output and p mode requirements with high purity limited only by the manufacturing processes and conversion efficiency.

The SPP has its own intrinsic set of azimuthal and radial mode numbers L and P respectively, which need not be integers. SPP’s [20] are simple in their construction in that the spatial phase pattern of the LG mode, namely the helical wavefront, is imprinted onto a glass substrate via a spiral increase in substrate thickness with the total step height H as shown in Fig. 1(a). The spiral height is proportional to the laser wavelength and material index given by the following: H = / (n − 1) [20–22]. Here, λ is the incident laser wavelength and n is the index of refraction of the SPP. It was reported [20], that the conversion efficiency of an L = 1, P = 0 spiral phase plate was 78.5% from the TEM00 mode to the first LG10 mode. It was later reported, that the conversion efficiency was higher yielding a conversion efficiency of around 93% to the LG10 mode [21]. The difference arises from the fact that the first report assumed an output beam waist equal to the input Gaussian beam waist while the second report allowed the output beam waist to be a free parameter. However, the second report does not give details about the relation of the output LG mode beam waist to the input Gaussian beam waist.

 figure: Fig. 1

Fig. 1 (a) Continuous spiral phase optic with L = 1 and P = 0 also indicating the total spiral height H. (b) 16 step spiral phase optic with L = 1 and P = 0. Both plates shown with an 800nm step height.

Download Full Size | PDF

Spiral phase mirrors (SPM’s) are an extension to this generation method using a reflective optical element [23,24] where the height of each spiral step is one half the input wavelength. Since their first publication, SPP’s and SPM’s have been manufactured using a variety of processes with deposition and erosion techniques. We will call these transmission and reflection optics collectively spiral phase optics (SPO’s).

The 2004 paper by Sueda et al. [25], introduced the idea of using a stepped spiral phase plate (SSPP) [Fig. 1(b)], rather than a continuous spiral [Fig. 1(a)] as first proposed [20]. By depositing steps using electron beam evaporation, fabrication of an SSPP is simpler with a small loss in conversion efficiency. It was reported [25] that the expected conversion efficiency to the LG10 mode dropped from a value of 78.5% to 77.5% when switching from a continuous phase plate to a 16 step L = 1, P = 0 spiral phase plate, again assuming the same output beam waist as input beam waist.

When considering the use of spiral phase optics, both continuous and stepped in ultrafast laser systems, the laser bandwidth can become a significant fraction of the central wavelength. Conversion efficiency calculations of SPO’s in laser beamlines have often assumed that the incident light is monochromatic, neglecting this effect. While a few reports have tried to address this issue [26], little work has been done in calculating the effects of broadband beams and spiral steps on mode generation purity and conversion efficiency. Here we show for the first time, an analytical expression for the conversion efficiency from a Gaussian TEM00 laser mode to an arbitrary Laguerre-Gaussian mode for continuous, stepped, and non-integer charged spiral phase optics for the lowest order radial component (P = 0). This analysis also includes an estimation of the effects from broad laser bandwidth using numerical techniques.

2. Mode decomposition

As a Gaussian TEM00 mode is transmitted through or reflected from a spiral phase optic, the spatial phase of the beam is shifted by the spatial phase pattern of the optic. Initially this mode in the near field remains spatially a phase shifted Gaussian TEM00 mode, but as it propagates to the far field the beam phase begins to reshape its intensity profile that can be described by a distribution of LG modes. The higher order p LG modes diverge more quickly than lower order modes leaving the lowest order p modes dominating in the far field near the axis. Conversion to each of these modes can be calculated from the inner product of the input and output modes with a transmission operator T, representative of the spatially varying phase shift from the optic. We can express the transmission operator for a continuous (infinite number of steps) SPO with P = 0 by the following expression:

T=exp(iLϕ)
Here we define L as the topological charge of the SPO, which dictates the fundamental azimuthal mode number of the output beam . In this paper we limit ourselves to a transmission operator that has a null radial index, that is, P = 0 as they are the most common in use today. Higher order P mode SPO’s will be discussed separately due to the mathematical complexities they introduce.

The general expression for an arbitrary normalized LG mode amplitude is given by [27]:

up=2p!π(p+||)!1w(z)[r2w(z)]||exp[r2w2(z)]Lp||(2r2w2(z))×exp[iϕ]exp[ik0r2z2(z2+zR2)]exp[iψ(z)]

Here, zR=πw02/λ is the Rayleigh range with wavelength λ and wavevector k0 = 2π/λ, w(z)=w0[(z2+zR2)/zR2]1/2 is the radius of the beam with beam waist w0 and ψ(z) = [(2p + || + 1) arctan (z/zR)] is the Gouy phase. The radial and azimuthal mode numbers of the LG mode, p and respectively, describe an array of unique modes where p is zero or any positive integer and is any real integer [28]. The associated Laguerre polynomials Lp can be given by the expression:

Lp(x)=n=0p(1)n(p+)!(pn)!(+n)!n!xn

The conversion efficiency from one orthogonal mode to another by means of a transfer function is generally given by the inner product [18]:

ηp=|up|T|umn|2=|02π0up*(r,ϕ,z)T(r,ϕ)umn(r,ϕ,z)rdrdϕ|2

The modes are normalized such that a unit transfer function in the inner product will result in 0 or 1, and squaring the magnitude of the inner product gives the normalized conversion efficiency amplitude ηℓp with transmission function T. Setting the mode numbers p = = 0 in Eq. (2), we obtain the familiar TEM00 mode common to most laser cavity outputs. Expanding the inner product in Eq. (4) with an input TEM00 mode and using a continuous spiral phase optic yields an integral with 2 free parameters, the beam waists and their axial positions z. In many previous works [18,20,25,29], the beam waists were assumed equal and analyzed with the waist positions both set to z = 0 to yield a unique mode decomposition. There is however, no physical requirement for the beam waist sizes to be equal [21, 22, 28]. Whilst the position of the beam waist may also vary, such a change leads to a modification of the Gouy phase for each beam and the integral becomes non-analytic. We therefore allow the beam waist to vary but maintain that the beam waists be set to z = 0. Using an input TEM00 mode with beam waist w0(z = 0), and choosing an output mode with mode numbers and p and beam waist w1(z = 0), the inner product Eq. (4) is given by:

up|T|u00=2πp!(p+||)!1w0w102π0[r2w1]||exp[r2(1w02+1w12)]×Lp||(2r2w12)exp[i(L)ϕ]rdrdϕ=R||pΦL

The double integral may be separated into the product of two integrals, denoted Rℓp and ΦℓL which represent the radial and azimuthal integrals respectively. Substituting the series expansion for the associated Laguerre polynomials, the radial part of the integral including the multiplicative coefficients becomes:

R||p=2πp!(p+||)!2||2w0w1||+1×0r||+1exp(βr2)m=0p(1)m(||+p)!(pm)!(||+m)!m!2mr2mw12mdr
With,
β=1w02+1w12

For the case where the SPO has a P = 0 radial index, one can find an analytic solution to the radial integral utilizing the following identity [30]:

0xnexp(axb)dx=1ba1/b(n+1)Γ(n+1b)
where Γ represents the gamma function. For the case of the continuous SPO, we can also evaluate the azimuthal integral below to yield 2π when the values of L and are equal:
ΦL=02πexp(i(L)ϕ)dϕ=2π|=L

When the difference between the topological charges of the SPO L and output mode is a nonzero integer, the integral evaluates to zero. As already stated, the azimuthal charge of the output beam must be an integer, but the topological charge of the SPO can be designed to have any value. Non-trivial solutions exist when the SPO carries a non-integer topological charge and will be discussed in more detail in the next section. Combining and simplifying the results from Eq. (6) and Eq. (9) yields the following result:

ηp=|R||pΦ=L|2=2||+2p!(||+p)!γ2[m=0p(2)mΓ(μ)(pm)!(||+m)!m!(1+γ2)μ]2

Here, we define the beam waist ratio, γ = w1/w0 and define μ = m + ||/2 + 1. This gives the general conversion efficiency of a TEM00 mode into any LG mode through an integer charged, continuous SPO with P = 0 and = L. The mode conversion efficiency can be thus calculated as a function of the beam waist ratio, the results of which are plotted in Fig. 2 for the case of L = = 1. The output from the SPO gives a spectrum of LG modes centered around a fundamental azimuthal mode (L = ) with corresponding radial modes p. It was shown [20] that a continuous spiral phase plate can convert a TEM00 mode into an LG10 mode with 78.5% maximum efficiency assuming the input and output beam waists are equal. However, the actual conversion efficiency into an LG10 mode is higher reaching just over 93% when the ratio of the beam waists is set to an optimum value. The maximum conversion efficiency to the fundamental mode can be found by finding the extremum of Eq. (10) where the slope is zero relative to γ while setting p = 0, giving the optimum value of γ as:

γ=1||+1

This expression gives the optimal beam waist ratio for conversion into any LG0 mode yielding a value of η10 = 0.9308 at γ=1/2 for a L = 1, P = 0 SPO. The remaining 7% of the input beam energy in this case is converted into higher order radial modes illustrated in Fig. 2. The conversion efficiency value of these additional modes is summarized in Table 1. An interesting observation made here is the conversion efficiency to odd p modes goes to zero at some value of the beam waist ratio with the p = 1 mode going to zero at the optimal beam waist ratio for all = L values.

 figure: Fig. 2

Fig. 2 Conversion efficiency of a continuous spiral phase optic with L = 1 to the first 5 LGℓp modes. The peak of the LG10 mode is located at γ=1/2 where γ is the ratio of the output beam waist to the input beam waist.

Download Full Size | PDF

Tables Icon

Table 1. Conversion efficiencies from a TEM00 mode to various LG modes for given SPO step numbers N. The SPO carries a null radial index (P = 0) and results are shown for both L = 1 and L = 2 optics. The N = ∞ corresponds to a continuous phase optic.

The reason for an optimum output beam waist parameter w1 which is smaller than the starting input beam waist parameter, w0, can be understood as follows. For a given beam waist parameter the root mean square (RMS) intensity radius of an LG10 mode is larger than the RMS intensity radius of a TEM00 mode [31]. Thus if one converts a TEM00 with a given RMS intensity radius into an LG10 mode with the same RMS intensity radius, the matching beam radius of the LG10 mode is necessarily smaller. For higher mode numbers the optimum matching radius becomes even smaller as shown in Eq. (11). An assumption made in this analysis which can experimentally lower this conversion efficiency is the domain of the radial integral, assumed to have an infinite boundary. Realistically, SPO’s are finite in radius, and therefore we assume that in practice the incident Gaussian spot size is small compared to the SPO diameter.

As one increases the topological charge of the SPO, the conversion efficiency to the fundamental = L mode decreases as shown in Fig. 3. The limit as L approaches infinity of Eq. (10) goes to zero. This implies that there is some limit to the efficient generation of OAM modes through the use of an SPO. If for example we require the conversion efficiency to the fundamental mode to be greater than 50%, we find that the maximum topological charge is L = 11. Likewise for a conversion efficiency of 20%, the charge on the plate can be no more than L = 82. The fabrication of high topological charge spiral phase mirrors (up to L = 100) have been reported previously [24], and have given interesting results. Conversion efficiencies to the fundamental mode were not however explicitly stated, it was stated however that the conversion efficiency was reduced in high charge plates due to scattering from the area between steps. Large topological charge modes have narrow ring-like intensity patterns limiting the amount of coupling from a broad Gaussian beam into such modes.

 figure: Fig. 3

Fig. 3 Conversion efficiency of a continuous spiral phase optic with charge L to the L = LG mode assuming the optimum output beam waist given by Eq. (11). Displayed are the results for the first four even modes: p = 0, 2, 4, 6 on a log-log scale.

Download Full Size | PDF

3. Stepped spiral phase optics

Extending the analysis to include stepped spiral phase optics requires modification of the transmission operator T, but not the radial integral R||p in the P = 0 case. In essence, stepped spiral phase optics (SSPO’s) are a spiral staircase consisting of N steps per revolution as shown in Fig. 1(b). Assuming all the steps are equally spaced and are of equal size, the transmission operator can be modified to match the SSPO phase in each step n given by:

T=exp[iL(2πnN)]

Again, the L value represents the intrinsic topological charge of the SSPO, additionally denoting n as the individual step number. Integrating each discrete step requires us to sum over all the steps to find the total contribution to the scaling factor. This can be written generally as the expression below.

ΦLN=n=0N12πn/N2π(n+1)/Nexp[i(ϕL2πnN)]dϕ
For the case where the SSPO and output mode azimuthal charges are equal ( = L), the integral can be evaluated and the sum converges to a finite value given by:
ΦLN=iNL[exp(iL2πN)1]

As one would expect, taking the limit of this expression when the number of steps approaches infinity yields 2π. This result is summarized in Fig. 4, and shows the conversion scaling factor |ΦLN|2 normalized to 4π2 as a function of N for various values of L. Setting N = 16 and γ = 1 in Eq. (14) returns a conversion efficiency of 77.5%, matching that in [25]. As expected, there is a minimum number of 2 steps required to convert to an LG10 mode, increasing as L increases. For the case of L = 1, a 16 step SSPO gives relative reduction in conversion efficiency of 1.28%. Similarly for an L = 2 plate, 32 steps gives the same relative reduction in conversion efficiency of 1.28%. These values are summarized in Table 1. It can be seen that an efficient SSPO (less than 1.3% conversion efficiency loss compared to a continuous plate), requires no more than 16L steps, for a topological charge L SSPO.

 figure: Fig. 4

Fig. 4 Multiplicative scaling factor |ΦLN|2 as a function of the number of SSPO steps N for the L = = 1, 2, 3 cases, normalized to 4π2. As the number of steps approaches ∞, all cases approach a value of 1.

Download Full Size | PDF

Returning to the case where the SSPO has a non-integer topological charge, the transmission operator remains unchanged, but the integral solution is different. Summing over each of the integral terms converges again to a finite value as shown below:

ΦLN=i[exp(i2πN)1]exp(i2π(L))1exp(i2π(L)/N)1
The limit as the total number of steps goes to infinity results in the solution:
limNΦLN=iL[exp(i2π(L))1]

This expression in Eq. (15) is quite general in that it includes both variables of the SSPO (N and L) and allows the use of an arbitrary output mode. As previously discussed, the output LG mode must contain an integer value azimuthal charge, however the design of the SPO can be such that L can be chosen to be any value. When L is a non-integer, additional azimuthal output modes i are also generated with the fundamental output charge . The fundamental charge is therefore defined as being the closest integer to L. The additional i modes generated produce a mixed state of LG modes whose net sum of angular momentum must equal the charge of the SPO. Conversion efficiency to the fundamental mode for a non-integer SPO is summarized in Fig. 5, where the mode conversion efficiency scaling factor |ΦℓLN|2, normalized to 4π2, is plotted as a function of the intrinsic charge of a continuous SPO converting to an LG10 output mode. The conversion efficiency to the fundamental mode is maximized when the SPO and output mode charges are equal, as expected. One sees that a fractional change in SPO charge decreases the conversion efficiency, dropping off significantly as the change approaches half an integer. However, for small mismatches of intrinsic charge due to step height and laser wavelength, the conversion efficiency remains close to its maximum value. From Fig. 5 it is clear that the efficiency change is less than 10% when the charge of the SPO is changed by a significant amount of 0.15. This is further summarized in Fig. 6 where we have plotted the spectral content of an output beam from a continuous L = 1.2 SPO for the case of γ=1/2. Here we see the = 1, p = 0 is still the fundamental mode as expected, with a conversion efficiency of 0.815, but adjacent i modes are also present, including the unconverted TEM00 mode.

 figure: Fig. 5

Fig. 5 Multiplicative scaling factor |ΦLN|2 as a function of intrinsic topological charge of the spiral phase optic L to the LG10 mode. Inset into the image is a wider field of view showing the behavior away from ( = L = 1), note the zeros of the function at all integer charges not equal to .

Download Full Size | PDF

4. Lasers with broad bandwidth

With this insight, we can investigate further by looking at the effect of a laser with bandwidth on the conversion efficiency, which in principle is the same as changing the charge of the SPO relative to the laser wavelength. Today, ultrafast lasers can carry a considerable bandwidth, pushing upwards of 200nm [32]. This spectrum can be roughly described by a Gaussian distribution in frequency centered around the laser frequency. To model this, we return to the phase transmission operator T for a SSPO, modified to include the laser central frequency ν0 and some arbitrary frequency value ν.

T=exp[iL(2πnνNν0)]

Again, this has no effect on the radial integral of Eq. (7) when examining the case of P = 0. If we convolve the conversion efficiency scaling factor |ΦℓLN|2 with a Gaussian distribution of frequencies, centered around ν0 we obtain the following:

|ΦLNν|2=1σ2πexp[12(νν0σ)2]×|n=0N12πn/N2π(n+1)/Nexp[i(ϕL2πnνNν0)]dϕ|2dν

We introduce the frequency standard deviation σ, which can be related to the FWMH (full width half maximum) of the laser bandwidth by the following identity: FWHM=22ln2σ. Evaluation of the azimuthal integral is still possible and leads to the result:

|ΦLNν|2=1σ2πexp[12(νν0σ)2]×|i[exp(i2πN)1]exp(i2π(Lνν0))1exp(i2π(Lνν0)/N)1|2dν

This general expression can be computed numerically. It may also be simplified further for specific cases of the SPO; continuous, integer SPO charge etc. The conversion efficiency scaling factor for a continuous SPO with equal charges (L = ) at the central wavelength is given in Fig. 7 for various representative wavelengths. From this figure, it is clear that when the laser bandwidth is equivalent to 10% of the laser wavelength, the conversion efficiency scaling factor is very close to 1. It should also be noted that as various spectral modes propagate to the far field, the different phase shifts of the overlapping spectral components may lead to complex spatial and temporal beam profiles.

 figure: Fig. 6

Fig. 6 Laguerre-Gauss mode spectrum (LGip) emanating from a continuous L = 1.2, P = 0 SPO. The azimuthal mode number i is on the x-axis, the radial mode number p is on the y-axis and the total conversion efficiency ηLℓp is displayed on the vertical z-axis on a log scale.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Conversion efficiency scaling factor ΦℓLNν as a function of laser bandwidth in nanometers for the case of = L and N = ∞. Four central wavelengths λ0 are plotted: 1064nm, 800nm, 532nm, 400nm.

Download Full Size | PDF

We have also neglected until now, effects of group velocity dispersion (GVD) in the SPP substrate. It is well understood that an ultrafast pulse will broaden when propagating in a material. This is not the case for SPM’s since there is no dispersion from an ideal reflecting surface, giving them an advantage in the ultrafast regime.

5. Conclusion

An analytical model for the conversion efficiency of a generic SPO with radial mode P = 0 was derived. A further analytic expression has been derived to include the effects of SSPO’s. Further integral expressions have been derived for the case of non-integer topological charge SPO’s and used to evaluate the effects of broad laser bandwidth. It was found that conversion from a TEM00 mode to the lowest order radial LG mode (p = 0) is most efficient when the ratio of output and input beam waists is chosen to be 1/||+1. The fraction of light not converted to this primary mode is converted into higher order radial modes which diverge more quickly than the fundamental mode yielding a beam profile dominated by the lowest order p mode in the far field. Additional i modes are generated for the non-integer and broadband laser cases. The fractional power converted into these modes is found to be low for small deviations in the SPO charge L as applicable to standard broad laser pulses.

It was also shown that as the topological charge L of the SPO is increased, the conversion efficiency decreases, independently of whether the SPO is smooth or stepped inferring a limit to the purity of the generated fundamental output mode.

In the case of SSPO’s, it was shown that using 16L steps results in less than a 1.3% relative reduction in conversion efficiency. The present results allow optimum design of SSPO’s, either transmissive or reflective for the conversion of a TEM00 mode to any fundamental OAM mode.

Funding

This work was funded by the Natural Sciences and Engineering Research Council of Canada (NSERC) research grant number RGPIN-2014-05736.

References and links

1. L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, and J. P. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A 45(11), 8185–8190 (1992). [CrossRef]   [PubMed]  

2. T. Kuga, Y. Torii, N. Shiokawa, T. Hirano, Y. Shimizu, and H. Sasada, “Novel optical trap of atoms with a doughnut beam,” Phys. Rev. Lett. 78, 4713–4716 (1997). [CrossRef]  

3. M. Padgett and R. Bowman, “Tweezers with a twist,” Nature Photon. 5, 343–348 (2011). [CrossRef]  

4. A. E. Willner, H. Huang, Y. Yan, Y. Ren, N. Ahmed, G. Xie, C. Bao, L. Li, Y. Cao, Z. Zhao, J. Wang, M. P. J. Lavery, M. Tur, S. Ramachandran, A. F. Molisch, N. Ashrafi, and S. Ashrafi, “Optical communications using orbital angular momentum beams,” Adv. Opt. Photon. 7(1), 66–106 (2015). [CrossRef]  

5. G. Molina Terriza, J. P. Torres, and L. Torner, “Twisted photons,” Nature Phys. 3, 305–310 (2007). [CrossRef]  

6. Y. Iketaki and N. Bokor, “Super-resolution microscope using a two-color phase plate for generating quasi-Laguerre-Gaussian beam,” Opt. Commun. 285(18), 3798–3804 (2012). [CrossRef]  

7. T. F. Scott, B. A. Kowalski, A. C. Sullivan, C. N. Bowman, and R. R. McLeod, “Two-color single-photon photoinitiation and photoinhibition for subdiffraction photolithography,” Science 324, 913–917 (2009). [CrossRef]   [PubMed]  

8. J. T. Mendonça, B. Thiede, and H. Then, “Stimulated Raman and Brillouin backscattering of collimated beams carrying orbital angular momentum,” Phys. Rev. Lett. 102, 185005 (2009). [CrossRef]   [PubMed]  

9. J. T. Mendonça, “Twisted waves in a plasma,” Plasma Phys. Control. Fusion 54, 124031 (2012). [CrossRef]  

10. J. Vieira and J. T. Mendonça, “Nonlinear laser driven donut wakefields for positron and electron acceleration,” Phys. Rev. Lett. 112, 215001 (2014). [CrossRef]  

11. C. Brabetz, S. Busold, T. Cowan, O. Deppert, D. Jahn, O. Kester, M. Roth, D. Schumacher, and V. Bagnoud, “Laser-driven ion acceleration with hollow laser beams,” Phys. Plasmas 22, 013105 (2015) [CrossRef]  

12. S. Ali, J. R. Davies, and J. T. Mendonça, “Inverse Faraday effect with linearly polarized laser pulses,” Phys. Rev. Lett. 105, 035001 (2010). [CrossRef]   [PubMed]  

13. T. V. Liseykina, S. V. Popruzhenko, and A. Macchi, “Inverse Faraday effect driven by radiation friction,” New J. Phys. 18, 072001 (2016). [CrossRef]  

14. Zs. Lecz, I. V. Konoplev, A. Seryi, and A. Andreev, “GigaGauss solenoidal magnetic field inside bubbles excited in under-dense plasma,” Sci. Rep. 6, 36139 (2016). [CrossRef]   [PubMed]  

15. G. B. Zhang, M. Chen, C. B. Schroeder, J. Luo, M. Zeng, F. Y. Li, L. L. Yu, S. M. Weng, Y. Y. Ma, T. P. Yu, Z. M. Sheng, and E. Esarey, “Acceleration and evolution of a hollow electron beam in wakefields driven by a Laguerre-Gaussian laser pulse,” Phys. Plasmas 23, 033114 (2016). [CrossRef]  

16. M. Zürch, C. Kern, P. Hansinger, A. Dreischuh, and Ch. Spielmann, “Strong-field physics with singular light beams,” Nature Phys. 8, 743–745 (2012). [CrossRef]  

17. A. Denoeud, L. Chopineau, A. Leblanc, and F. Quéré, “Interaction of ultraintense laser vortices with plasma mirrors,” Phys. Rev. Lett. 118, 033902 (2017). [CrossRef]   [PubMed]  

18. M. W. Beijersbergen, L. Allen, H.E.L.O. van der Veen, and J. P. Woerdman, “Astigmatic laser mode converotrs and transfer of orbital angular momentum,” Opt. Commun. 96(1,2,3), 123–132 (1993). [CrossRef]  

19. J. Strohaber, T. D. Scarborough, and C. J. G. J. Uiterwaal, “Ultrashort intense-field optical vortices produced wit laser-etched mirrors,” Appl. Opt. 46(36), 8583–8590 (2007). [CrossRef]   [PubMed]  

20. M. W. Beijersbergen, R. P. C. Coerwinkel, M. Kristensen, and J. P. Woerdman, “Helical-wavefront laser beams produced with a spiral phaseplate,” Opt. Commun. 112, 321–327 (1994). [CrossRef]  

21. G. Ruffato, M. Massari, and F. Romanato, “Generation of high-order Laguerre-Gaussian modes by means of spiral phase plates,” Opt. Lett. 39, 5094–5097 (2014). [CrossRef]   [PubMed]  

22. G. Ruffato, M. Massari, M. Carli, and F. Romanato, “Spiral phase plates with radial discontinuities for the generation of multiring orbital angular momentum beams: fabrication, characterization, and application,” Opt. Eng. 54(11), 111307 (2015). [CrossRef]  

23. D. P. Ghai, “Generation of optical vortices with an adaptive helical mirror,” Appl. Opt. 50(10), 1374–1381 (2011). [CrossRef]   [PubMed]  

24. G. Campbell, B. Hage, B. Buchler, and P. K. Lam, “Generation of high-order optical vortices using directly machined spiral phase mirrors,” Appl. Opt. 51(7), 873–876 (2012). [CrossRef]   [PubMed]  

25. K. Sueda, G. Miyaji, N. Miyanaga, and M. Nakatsuka, “Laguerre-Gaussian beam generated with a multilevel spiral phase plate for high intensity laser pulses,” Opt. Express 12(15), 3548–3553 (2004) [CrossRef]   [PubMed]  

26. Q. Xie and D. Zhao, “Optical vortices generated by multi-level achromatic spiral phase plates for broadband beams,” Opt. Commun. 281, 7–11 (2008). [CrossRef]  

27. A. M. Yao and M. J. Padgett, “Orbital angular momentum: origins, behavior and applications,” Adv. Opt. Photon. 3, 161–204 (2011). [CrossRef]  

28. A. E. Siegman, Lasers (University Science Books, 1986).

29. Y. Shi, B. Shen, L. Zhang, X. Zhang, W. Wang, and Z. Xu, “Light fan driven by a relativistic laser pulse,” Phys. Rev. Lett. 112, 235001 (2014). [CrossRef]   [PubMed]  

30. I. Gradshteyn and I. Ryzhik, Table of Integrals, Series, and Products, 7th ed. (Elsevier, 2007).

31. M. J. Padgett and L. Allen, “The Poynting vector in Laguerre-Gaussian laser modes,” Opt. Commun. 121, 36–40 (1995). [CrossRef]  

32. S. Backus, C. G. Durfee, M. M. Murname, and H. C. Kapteyn, “High power ultrafast lasers,” Rev. Sci. Instrum. 69(3), 1207–1223 (1998). [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 (7)

Fig. 1
Fig. 1 (a) Continuous spiral phase optic with L = 1 and P = 0 also indicating the total spiral height H. (b) 16 step spiral phase optic with L = 1 and P = 0. Both plates shown with an 800nm step height.
Fig. 2
Fig. 2 Conversion efficiency of a continuous spiral phase optic with L = 1 to the first 5 LGℓp modes. The peak of the LG10 mode is located at γ = 1 / 2 where γ is the ratio of the output beam waist to the input beam waist.
Fig. 3
Fig. 3 Conversion efficiency of a continuous spiral phase optic with charge L to the L = LG mode assuming the optimum output beam waist given by Eq. (11). Displayed are the results for the first four even modes: p = 0, 2, 4, 6 on a log-log scale.
Fig. 4
Fig. 4 Multiplicative scaling factor |ΦLN|2 as a function of the number of SSPO steps N for the L = = 1, 2, 3 cases, normalized to 4π2. As the number of steps approaches ∞, all cases approach a value of 1.
Fig. 5
Fig. 5 Multiplicative scaling factor |ΦLN|2 as a function of intrinsic topological charge of the spiral phase optic L to the LG10 mode. Inset into the image is a wider field of view showing the behavior away from ( = L = 1), note the zeros of the function at all integer charges not equal to .
Fig. 6
Fig. 6 Laguerre-Gauss mode spectrum (LGip) emanating from a continuous L = 1.2, P = 0 SPO. The azimuthal mode number i is on the x-axis, the radial mode number p is on the y-axis and the total conversion efficiency ηLℓp is displayed on the vertical z-axis on a log scale.
Fig. 7
Fig. 7 Conversion efficiency scaling factor ΦℓLNν as a function of laser bandwidth in nanometers for the case of = L and N = ∞. Four central wavelengths λ0 are plotted: 1064nm, 800nm, 532nm, 400nm.

Tables (1)

Tables Icon

Table 1 Conversion efficiencies from a TEM00 mode to various LG modes for given SPO step numbers N. The SPO carries a null radial index (P = 0) and results are shown for both L = 1 and L = 2 optics. The N = ∞ corresponds to a continuous phase optic.

Equations (19)

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

T = exp ( i L ϕ )
u p = 2 p ! π ( p + | | ) ! 1 w ( z ) [ r 2 w ( z ) ] | | exp [ r 2 w 2 ( z ) ] L p | | ( 2 r 2 w 2 ( z ) ) × exp [ i ϕ ] exp [ i k 0 r 2 z 2 ( z 2 + z R 2 ) ] exp [ i ψ ( z ) ]
L p ( x ) = n = 0 p ( 1 ) n ( p + ) ! ( p n ) ! ( + n ) ! n ! x n
η p = | u p | T | u m n | 2 = | 0 2 π 0 u p * ( r , ϕ , z ) T ( r , ϕ ) u m n ( r , ϕ , z ) r d r d ϕ | 2
u p | T | u 00 = 2 π p ! ( p + | | ) ! 1 w 0 w 1 0 2 π 0 [ r 2 w 1 ] | | exp [ r 2 ( 1 w 0 2 + 1 w 1 2 ) ] × L p | | ( 2 r 2 w 1 2 ) exp [ i ( L ) ϕ ] r d r d ϕ = R | | p Φ L
R | | p = 2 π p ! ( p + | | ) ! 2 | | 2 w 0 w 1 | | + 1 × 0 r | | + 1 exp ( β r 2 ) m = 0 p ( 1 ) m ( | | + p ) ! ( p m ) ! ( | | + m ) ! m ! 2 m r 2 m w 1 2 m d r
β = 1 w 0 2 + 1 w 1 2
0 x n exp ( a x b ) d x = 1 b a 1 / b ( n + 1 ) Γ ( n + 1 b )
Φ L = 0 2 π exp ( i ( L ) ϕ ) d ϕ = 2 π | = L
η p = | R | | p Φ = L | 2 = 2 | | + 2 p ! ( | | + p ) ! γ 2 [ m = 0 p ( 2 ) m Γ ( μ ) ( p m ) ! ( | | + m ) ! m ! ( 1 + γ 2 ) μ ] 2
γ = 1 | | + 1
T = exp [ i L ( 2 π n N ) ]
Φ L N = n = 0 N 1 2 π n / N 2 π ( n + 1 ) / N exp [ i ( ϕ L 2 π n N ) ] d ϕ
Φ L N = i N L [ exp ( i L 2 π N ) 1 ]
Φ L N = i [ exp ( i 2 π N ) 1 ] exp ( i 2 π ( L ) ) 1 exp ( i 2 π ( L ) / N ) 1
lim N Φ L N = i L [ exp ( i 2 π ( L ) ) 1 ]
T = exp [ i L ( 2 π n ν N ν 0 ) ]
| Φ L N ν | 2 = 1 σ 2 π exp [ 1 2 ( ν ν 0 σ ) 2 ] × | n = 0 N 1 2 π n / N 2 π ( n + 1 ) / N exp [ i ( ϕ L 2 π n ν N ν 0 ) ] d ϕ | 2 d ν
| Φ L N ν | 2 = 1 σ 2 π exp [ 1 2 ( ν ν 0 σ ) 2 ] × | i [ exp ( i 2 π N ) 1 ] exp ( i 2 π ( L ν ν 0 ) ) 1 exp ( i 2 π ( L ν ν 0 ) / N ) 1 | 2 d ν
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.