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

Super amplification enabled by orbital angular momentum in weak measurement

Open Access Open Access

Abstract

Weak measurement, which can amplify a weak signal, has shown great significance in precision measurements. The amplification is usually realized through the weak value and the propagation factor. We show that the orbital angular momentum (OAM) can provide another dimension for amplification that is linearly proportional to the OAM number. We employ OAM to measure the spin Hall effect of light and demonstrate that the OAM-enabled amplification is compatible with the weak value amplification and the propagation amplification. This work is probable to promote the application of OAM in precision measurements.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The concept of weak measurement was originally proposed by Aharonov, Albert, and Vaidman [1]. It was pointed out that with appropriate pre- and post-selections in a quantum measurement, the eigenvalue of a quantum state can be amplified by the weak value, which is now known as the weak value amplification (WVA) technique. The intriguing property of weak measurement has attracted widespread discussion and sparked many novel thoughts in quantum physics [28]. The interpretation of the weak value still remains an open question [913]. In a practical sense, weak measurement has also shown great power in precision parameter estimations since a tiny signal can be greatly amplified to be detectable. For example, the position and momentum displacements [1416], the frequency and time shifts [1720], the phase difference [21,22], and even the angular rotations [23] have effectively been amplified by the weak value.

In 2008, Hosten and Kwiat found that apart from the weak value, a propagation factor could be arranged to realize an additional amplification in weak measurement [24], such that the spin Hall effect of light, which results in a spatial displacement on the scale of nanometers (far below the pixel resolution of a scientific camera), can be successfully observed. From then on, the propagation amplification (PA) has been jointly applied in many precision measurement scenarios. Reports have shown that an angular deflection of a mirror down to 400 frad [25], an optical rotation angle of 0.2 mdeg [26], and a phase difference on the order of $10^{-7}$ rad [27] can be measured under the WVA and the PA.

In this work, we take into account the orbital angular momentum (OAM) that could accommodate a new dimension for amplification. Several studies have preliminarily demonstrated the feasibility of applying OAM to achieve an amplification in weak measurements, which, however, neglected the interaction of OAM with other amplifications [2831]. We use the vortex beam that carries OAM to measure the spin Hall effect of light which can lead to many practically meaningful applications [26,32,33]. The results provide a compelling evidence that the OAM amplification can work together with the WVA and the PA.

2. Theoretical analysis

The experimental setup is shown in Fig. 1. We use a He-Ne laser with the wavelength of $\lambda =633$ nm as the Gaussian light source. The first polarizer (P1) makes the beam in the horizontal polarization state, which matches the working axis of the following phase-only spatial light modulator (SLM). The half wave plate (HWP1) can be used to adjust the light intensity. We load a fork-shaped grating in the SLM, such that a vortex beam is transformed in the first diffraction order. A 4$f$ system, which is made up of two lenses (L1 and L2) with the focal lengths both being 100 mm and an iris diaphragm, is built to filter out the first diffraction order. The wave function of the generated vortex beam can be written as [34]

$$\psi_0(x,y)\propto(x+is_my)^{|m|}e^{-\frac{x^2+y^2}{2w^2}},$$
where $m$ denotes the OAM number, $s_m=\text {sign}(m)$ and $w$ denotes the waist radius. Here and hereafter, we use the proportionality sign $"\propto "$ instead of the equality sign $"="$ in wave functions to omit inessential factors.

 figure: Fig. 1.

Fig. 1. Schematic diagram of the experimental setup. We use a He-Ne laser as the light source, a half wave plate (HWP1), three polarizers (P1, P2 and P3), a phase-only spatial light modulator (SLM), four lenses (L1, L2, L3 and L4), an Iris diaphragm (Iris), a prism, and a charge coupled device (CCD).

Download Full Size | PDF

The second polarizer (P2) is placed to pre-select the polarization state in

$$|{i}\rangle=\frac{1}{\sqrt{2}}(|{L}\rangle+|{R}\rangle),$$
where $|{L}\rangle$ and $|{R}\rangle$ represent the left-handed and right-handed circular polarization states, respectively. Lenses 3 and 4 (L3 and L4) construct a telescope system, and a prism is inserted between the two lenses. We denote the focal lengths of L3 and L4 by $f_1$ and $f_2$, respectively. The distance between L3 and the prism is $f_1$ and accordingly, the distance between the prism and L4 is $f_2$. It is well known that the wave function in the back focal plane of a lens is given by the Fourier transform of the wave function in the front focal plane. Therefore, we can immediately obtain the wave function in the back focal plane of L3 as
$$\psi_1(x,y)\propto(x+is_my)^{|m|}e^{-\frac{x^2+y^2}{2(\lambda f_1/2\pi w)^2}},$$
which indicates that the light beam focused by a lens still carries OAM, but the waist is changed from $w$ to $\lambda f_1/2\pi w$.

When the light beam is reflected by the prism, the component in the state $|{L}\rangle$ will be shifted along the transversal direction ($y$-axis) by a value of $g$, and the component in $|{R}\rangle$ will be shifted along the opposite direction by the same value. This effect is termed the spin Hall effect of light, which is an analogue of the Hall effect. In the Hall effect, if a perpendicular magnetic field is applied to a conductor, the positive and negative charges are accumulated in opposite surfaces. While in the spin Hall effect of light, the left-handed and right-handed circular polarization components are spatially separated due to the refractive index gradient on the surface of the prism. It can be derived that $g=-\frac {\cot \theta }{k}(1+\frac {r_p}{r_s})$, where $\theta$ is the incident angle, $k=2\pi /\lambda$ is the wave number, $r_p$ and $r_s$ are the Fresnel coefficients of the prism [35]. In our experiment, the incident angle is set to $\theta \approx 30^\circ$ and the prism is fabricated from N-BK7. So it follows that $r_s\approx -0.246$, $r_p\approx 0.164$ and thereby, $g\approx 58$ nm.

In the back focal plane of L4, we have the wave function that is obtained by the Fourier transform of the reflected light field on the prism, which indicates that the spatial shift will be transformed to an angular shift of the value $\gamma =kg/f_2$. The anisotropic angular shifts can be well represented by the unitary transformation

$$U=e^{{-}i\gamma\hat{A}\hat{y}},$$
where the Pauli spin operator $\hat {A}=|{L}\rangle\langle{L}|-|{R}\rangle\langle{R}|$ and $\hat {y}$ is the position operator. The unitary transformation indicates a coupling between the polarization state and the position space, with $\gamma$ being regarded as the coupling strength. We may assume that $\gamma$ is sufficiently small, such that $U$ can be expanded and kept only to the first order, i.e., $U\approx 1-i \gamma \hat {A}\hat {y}$, which is recognized as the weak coupling condition. Another polarizer (P3), with the polarizing axis being nearly orthogonal to that of P2, plays the role of post-selection and the post-selected state takes the form of
$$|{f}\rangle=\frac{1}{\sqrt{2}}(e^{i\alpha}|{L}\rangle-e^{{-}i\alpha}|{R}\rangle),$$
where $\alpha$ corresponds to the rotation angle from $90^\circ$ of P3. Then the weak value is obtained by
$$A_w=\frac{\langle{f}|{A}|{i}\rangle}{\langle{f}|{i}\rangle} =i\cot\alpha.$$

Under these circumstances, it follows the wave function in the back focal plane of L4 that

$$\begin{aligned} \psi_2(x,y) & \propto \langle{f}|e^{{-}i\gamma\hat{A}\hat{y}}(x+is_my)^{|m|}e^{-\frac{x^2+y^2}{2w^2(f_2/f_1)^2}}|{i}\rangle \\ & \propto(1-i\gamma A_wy)(x+is_my)^{|m|}e^{-\frac{x^2+y^2}{2w^2(f_2/f_1)^2}} \\ & =(1+\gamma\cot\alpha y)(x+is_my)^{|m|}e^{-\frac{x^2+y^2}{2w^2(f_2/f_1)^2}}. \end{aligned}$$

We utilize a charge coupled device (CCD) to detect the intensity distribution, which is given by

$$I(x,y)=|\psi_2(x,y)|^2\propto(1+2\gamma\cot\alpha y+\gamma^2\cot^2\alpha y^2)(x^2+y^2)^{|m|}e^{-\frac{x^2+y^2}{w^2(f_2/f_1)^2}}.$$

It can be found that the change of $\gamma$ is directly reflected in the change of $I(x,y)$ along the $y$ direction. Therefore, it is convenient to choose the centroid of light spot as the pointer variable, the explicit expression of which reads

$$\bar{y}=\frac{\iint_{-\infty}^{+\infty}yI(x,y)dxdy}{\iint_{-\infty}^{+\infty}I(x,y)dxdy}=\frac{2(|m|+1)\gamma w^2\cot\alpha(\frac{f_2}{f_1})^2}{(|m|+1)\gamma^2w^2\cot^2\alpha(\frac{f_2}{f_1})^2+2}.$$

3. Experimental results

In the experiment, we initially set $f_1=50$ mm and $f_2=100$ mm. By rotating P3 to make $\alpha$ ranging from $-4.8^\circ$ to $4.8^\circ$, we record each $I(x,y)$ and then calculate $\bar {y}$. As the OAM number is respectively taken $m=1$ and $m=3$, the results are depicted in Fig. 2. The curves and the dots are corresponding to the theoretical expectations and the experimental data, which exhibit a good agreement. It can be seen that the centroid shifts of $m=3$ are larger than that of $m=1$, implying an amplification brought by OAM. However, when $\alpha$ approaches zero, the centroid shifts of $m=1$ and $m=3$ are identical, implying that $\bar {y}$ is independent of $m$. This could be readily account for if we look at Eq. (9). Consider that $\alpha \rightarrow 0$, $\cot \alpha$ becomes extremely large such that $(|m|+1)\gamma ^2w^2\cot ^2\alpha (\frac {f_2}{f_1})^2\gg 2$. Equation. (9) is simplified to $\bar {y}\approx 2\alpha /\gamma$, which indicates that $\bar {y}$ is inversely proportional to $\gamma$ and there is no amplification of $\gamma$. Therefore, we should make $\alpha$ small and maintain the relation $(|m|+1)\gamma ^2w^2\cot ^2\alpha (\frac {f_2}{f_1})^2\ll 2$ in the meantime. It immediately follows that

$$\bar{y}\approx(|m|+1)\gamma w^2\cot\alpha(\frac{f_2}{f_1})^2=g(|m|+1)\frac{kw^2f_2}{f_1^2}\cot\alpha.$$

Consequently, the spatial shift $g$, which is due to the spin Hall effect of light, is amplified by three factors, the OAM number $|m|$, the so-called propagation factor $kw^2f_2/f_1^2$, and the imaginary weak value $A_w=i\cot \alpha$. If the input light field $\psi _0(x,y)$ is just a Gaussian wave function ($m=0$), then Eq. (10) will degrade into the standard form that has been widely applied in weak measurements.

 figure: Fig. 2.

Fig. 2. The centroid shift $\bar {y}$ in a function of the post-selected angle $\alpha$. The red dots and blue dots are corresponding to experimental results when the OAM number $m=1$ and $m=3$, respectively. The red and blue fitting curves indicate the theoretical expectation. Three intensity distributions are inset.

Download Full Size | PDF

To intuitively demonstrate the amplification, we measure the centroid shifts with respect to the OAM number ranging from 0 to 5. Figure 3 shows the experimental results when the post-selected angle $\alpha$ is taken $3^\circ$ and $4^\circ$, respectively. For each OAM number, we use a CCD to record 10 images of the intensity distribution, with the time interval of 1 second. Then the averages and the standard deviations of $\bar {y}$, which correspond to the dots and the error bars, can be calculated. The fitting lines is plotted based on the theoretical expectation Eq. (10). It can be found that the experimental results of $\bar {y}$ are larger than the expectations when $m=0$ and $m=5$. Whereas the experimental results of $m=1, 2, 3,$ and 4 are all slightly smaller. We may infer that the discrepancies between experiment and theory mainly come from the modulation error of SLM. The gradient of line increases as the post-selected angle reduces. The ratio of the two gradients can be used to compare the amplifications when two different post-selected angles are chosen. We can obtain that the ratio of $\alpha =3^\circ$ and $\alpha =4^\circ$ reads 1.27, which is close to the theoretical value $\cot 3^\circ /\cot 4^\circ \approx 1.33$.

 figure: Fig. 3.

Fig. 3. The centroid shift $\bar {y}$ in a function of the OAM number $m$. The magenta and purple dots are corresponding to the experimental results when the post-selected angle $\alpha =0.0419$ rad and $\alpha =0.0559$ rad, respectively. The fitting lines indicate the theoretical expectation.

Download Full Size | PDF

We further investigate the effect of propagation factor on the centroid shift. The post-selected angle is fixed at $3^\circ$, and the focal length of L4 is manually made to be $f_2=100$ mm and $f_2=150$ mm, respectively. For each OAM number, we also record 10 images of the intensity distribution by setting the time interval to be 1 second. Then we plot the experimental results in Fig. 4 with dots. It can be obtained that the ratio of the two gradients corresponding to $f_2=100$ mm and $f_2=150$ mm is 1.52 in experiment, which agrees well with the theoretical expectation $f_2/f_1=1.50$. It is worthy of noting that similar to the results in Fig. 3, the experimentally measured values of $\bar {y}$ are all larger than that in the fitting line when $m=0$ and $m=5$, but become smaller when $m=1, 2, 3,$ and 4. This feature again confirms that the discrepancies primarily result from the modulation error of SLM.

 figure: Fig. 4.

Fig. 4. The centroid shift $\bar {y}$ in a function of the OAM number $m$. The green and blue dots are corresponding to the experimental results when the focal length of L4 $f_2=100$ mm and $f_2=150$ mm, respectively. The fitting lines indicate the theoretical expectation.

Download Full Size | PDF

It is known that the precision in measurement is ultimately limited by the shot noise of detector, which scales as $\sqrt {N_d}$ with $N_d$ denoting the photon number received by detector. In our scheme, it follows that $N_d=|\langle {f}|{i}\rangle |^2N$ as $N$ photons are initially incident. So the signal-to-noise ratio (SNR) is proportional to $(|m|+1)/\sqrt {|\langle {f}|{i}\rangle |^2N}$, which means that the OAM amplification can also increase the SNR.

4. Conclusion

In summary, we have shown that OAM can be utilized to achieve a super amplification aside from the WVA and the PA in weak measurement. We have proposed a scheme to measure the spatial displacement that is induced by the spin Hall effect of light. The experimental results showed that the centroid shift can be not only amplified by the weak value and propagation factor, but also proportional to the OAM number. OAM has the ability to offer another dimension for amplifying a weak signal, which could find important applications in a variety of precision measurement scenarios.

Funding

the Yibin University high level talents sailing plan project (No. 2021QH05); the Open Research Fund of Computational Physics Key Laboratory of Sichuan Province, Yibin University (No. 412-2020JSWLZD009).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. Y. Aharonov, D. Z. Albert, and L. Vaidman, “How the result of a measurement of a component of the spin of a spin-1/2 particle can turn out to be 100,” Phys. Rev. Lett. 60(14), 1351–1354 (1988). [CrossRef]  

2. J. S. Lundeen and A. M. Steinberg, “Experimental joint weak measurement on a photon pair as a probe of hardy’s paradox,” Phys. Rev. Lett. 102(2), 020404 (2009). [CrossRef]  

3. K. Yokota, T. Yamamoto, M. Koashi, et al., “Direct observation of hardy’s paradox by joint weak measurement with an entangled photon pair,” New J. Phys. 11(3), 033011 (2009). [CrossRef]  

4. Y. Kim, D.-G. Im, Y.-S. Kim, et al., “Observing the quantum cheshire cat effect with noninvasive weak measurement,” npj Quantum Inf. 7(1), 13 (2021). [CrossRef]  

5. T. Denkmayr, H. Geppert, S. Sponar, et al., “Observation of a quantum cheshire cat in a matter-wave interferometer experiment,” Nat. Commun. 5(1), 4492 (2014). [CrossRef]  

6. A. Romito, Y. Gefen, and Y. M. Blanter, “Weak values of electron spin in a double quantum dot,” Phys. Rev. Lett. 100(5), 056801 (2008). [CrossRef]  

7. N. Brunner, V. Scarani, M. Wegmüller, et al., “Direct measurement of superluminal group velocity and signal velocity in an optical fiber,” Phys. Rev. Lett. 93(20), 203902 (2004). [CrossRef]  

8. D. R. Solli, C. F. McCormick, R. Y. Chiao, et al., “Fast light, slow light, and phase singularities: A connection to generalized weak values,” Phys. Rev. Lett. 92(4), 043601 (2004). [CrossRef]  

9. J. Dressel and A. N. Jordan, “Significance of the imaginary part of the weak value,” Phys. Rev. A 85(1), 012107 (2012). [CrossRef]  

10. J. Dressel, M. Malik, F. M. Miatto, et al., “Colloquium: Understanding quantum weak values: Basics and applications,” Rev. Mod. Phys. 86(1), 307–316 (2014). [CrossRef]  

11. C. Ferrie and J. Combes, “Weak value amplification is suboptimal for estimation and detection,” Phys. Rev. Lett. 112(4), 040406 (2014). [CrossRef]  

12. A. N. Jordan, J. Martínez-Rincón, and J. C. Howell, “Technical advantages for weak-value amplification: When less is more,” Phys. Rev. X 4, 011031 (2014). [CrossRef]  

13. G. C. Knee and E. M. Gauger, “When amplification with weak values fails to suppress technical noise,” Phys. Rev. X 4, 011032 (2014). [CrossRef]  

14. N. W. M. Ritchie, J. G. Story, and R. G. Hulet, “Realization of a measurement of a “weak value”,” Phys. Rev. Lett. 66(9), 1107–1110 (1991). [CrossRef]  

15. R. Wang, J. Zhou, K. Zeng, et al., “Ultrasensitive and real-time detection of chemical reaction rate based on the photonic spin hall effect,” APL Photonics 5(1), 016105 (2020). [CrossRef]  

16. B. Xia, J. Huang, C. Fang, et al., “High-precision multiparameter weak measurement with hermite-gaussian pointer,” Phys. Rev. Appl. 13(3), 034023 (2020). [CrossRef]  

17. N. Brunner and C. Simon, “Measuring small longitudinal phase shifts: Weak measurements or standard interferometry?” Phys. Rev. Lett. 105(1), 010405 (2010). [CrossRef]  

18. X.-Y. Xu, Y. Kedem, K. Sun, et al., “Phase estimation with weak measurement using a white light source,” Phys. Rev. Lett. 111(3), 033604 (2013). [CrossRef]  

19. Y. Xu, L. Shi, T. Guan, et al., “Multifunctional weak measurement system that can measure the refractive index and optical rotation of a solution,” Appl. Phys. Lett. 114(18), 181901 (2019). [CrossRef]  

20. G. I. Viza, J. Martínez-Rincón, G. A. Howland, et al., “Weak-values technique for velocity measurements,” Opt. Lett. 38(16), 2949–2952 (2013). [CrossRef]  

21. L. Li, Y. Li, Y.-L. Zhang, et al., “Phase amplification in optical interferometry with weak measurement,” Phys. Rev. A 97(3), 033851 (2018). [CrossRef]  

22. N. Modak, A. B. S. A. K. Singh, and N. Ghosh, “Generalized framework of weak-value amplification in path interference of polarized light for the enhancement of all possible polarization anisotropy effects,” Phys. Rev. A 103(5), 053518 (2021). [CrossRef]  

23. O. S. Magaña Loaiza, M. Mirhosseini, B. Rodenburg, et al., “Amplification of angular rotations using weak measurements,” Phys. Rev. Lett. 112(20), 200401 (2014). [CrossRef]  

24. O. Hosten and P. Kwiat, “Observation of the spin hall effect of light via weak measurements,” Science 319(5864), 787–790 (2008). [CrossRef]  

25. P. B. Dixon, D. J. Starling, A. N. Jordan, et al., “Ultrasensitive beam deflection measurement via interferometric weak value amplification,” Phys. Rev. Lett. 102(17), 173601 (2009). [CrossRef]  

26. X. Qiu, L. Xie, X. Liu, et al., “Estimation of optical rotation of chiral molecules with weak measurements,” Opt. Lett. 41(17), 4032–4035 (2016). [CrossRef]  

27. L. Xu, L. Luo, H. Wu, et al., “Measurement of chiral molecular parameters based on a combination of surface plasmon resonance and weak value amplification,” ACS Sens. 5(8), 2398–2407 (2020). [CrossRef]  

28. G. Puentes, N. Hermosa, and J. P. Torres, “Weak measurements with orbital-angular-momentum pointer states,” Phys. Rev. Lett. 109(4), 040401 (2012). [CrossRef]  

29. J. Qiu, C. Ren, and Z. Zhang, “Precisely measuring the orbital angular momentum of beams via weak measurement,” Phys. Rev. A 93(6), 063841 (2016). [CrossRef]  

30. J. Zhu, P. Zhang, Q. Li, et al., “Measuring the topological charge of orbital angular momentum beams by utilizing weak measurement principle,” Sci. Rep. 9(1), 7993 (2019). [CrossRef]  

31. W. Long, J. Pan, X. Guo, et al., “Optimized weak measurement of orbital angular momentum-induced beam shifts in optical reflection,” Photonics Res. 7(11), 1273–1278 (2019). [CrossRef]  

32. X. Qiu, X. Zhou, D. Hu, et al., “Determination of magneto-optical constant of Fe films with weak measurements,” Appl. Phys. Lett. 105(13), 131111 (2014). [CrossRef]  

33. L. Luo, X. Qiu, L. Xie, et al., “Precision improvement of surface plasmon resonance sensors based on weak-value amplification,” Opt. Express 25(18), 21107–21114 (2017). [CrossRef]  

34. M. Merano, N. Hermosa, J. P. Woerdman, et al., “How orbital angular momentum affects beam shifts in optical reflection,” Phys. Rev. A 82(2), 023817 (2010). [CrossRef]  

35. K. Y. Bliokh and A. Aiello, “Goos–hänchen and imbert–fedorov beam shifts: an overview,” J. Opt. 15(1), 014001 (2013). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (4)

Fig. 1.
Fig. 1. Schematic diagram of the experimental setup. We use a He-Ne laser as the light source, a half wave plate (HWP1), three polarizers (P1, P2 and P3), a phase-only spatial light modulator (SLM), four lenses (L1, L2, L3 and L4), an Iris diaphragm (Iris), a prism, and a charge coupled device (CCD).
Fig. 2.
Fig. 2. The centroid shift $\bar {y}$ in a function of the post-selected angle $\alpha$. The red dots and blue dots are corresponding to experimental results when the OAM number $m=1$ and $m=3$, respectively. The red and blue fitting curves indicate the theoretical expectation. Three intensity distributions are inset.
Fig. 3.
Fig. 3. The centroid shift $\bar {y}$ in a function of the OAM number $m$. The magenta and purple dots are corresponding to the experimental results when the post-selected angle $\alpha =0.0419$ rad and $\alpha =0.0559$ rad, respectively. The fitting lines indicate the theoretical expectation.
Fig. 4.
Fig. 4. The centroid shift $\bar {y}$ in a function of the OAM number $m$. The green and blue dots are corresponding to the experimental results when the focal length of L4 $f_2=100$ mm and $f_2=150$ mm, respectively. The fitting lines indicate the theoretical expectation.

Equations (10)

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

ψ 0 ( x , y ) ( x + i s m y ) | m | e x 2 + y 2 2 w 2 ,
| i = 1 2 ( | L + | R ) ,
ψ 1 ( x , y ) ( x + i s m y ) | m | e x 2 + y 2 2 ( λ f 1 / 2 π w ) 2 ,
U = e i γ A ^ y ^ ,
| f = 1 2 ( e i α | L e i α | R ) ,
A w = f | A | i f | i = i cot α .
ψ 2 ( x , y ) f | e i γ A ^ y ^ ( x + i s m y ) | m | e x 2 + y 2 2 w 2 ( f 2 / f 1 ) 2 | i ( 1 i γ A w y ) ( x + i s m y ) | m | e x 2 + y 2 2 w 2 ( f 2 / f 1 ) 2 = ( 1 + γ cot α y ) ( x + i s m y ) | m | e x 2 + y 2 2 w 2 ( f 2 / f 1 ) 2 .
I ( x , y ) = | ψ 2 ( x , y ) | 2 ( 1 + 2 γ cot α y + γ 2 cot 2 α y 2 ) ( x 2 + y 2 ) | m | e x 2 + y 2 w 2 ( f 2 / f 1 ) 2 .
y ¯ = + y I ( x , y ) d x d y + I ( x , y ) d x d y = 2 ( | m | + 1 ) γ w 2 cot α ( f 2 f 1 ) 2 ( | m | + 1 ) γ 2 w 2 cot 2 α ( f 2 f 1 ) 2 + 2 .
y ¯ ( | m | + 1 ) γ w 2 cot α ( f 2 f 1 ) 2 = g ( | m | + 1 ) k w 2 f 2 f 1 2 cot α .
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.