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

Performance characterization of scanning beam steered by tilting double prisms

Open Access Open Access

Abstract

A pair of orthogonal tilting double prisms with a tracking precision better than submicroradian order exhibits a good application potential in laser tracking fields. In the paper, the beam scanning performance determined by both the structure parameters and the tilting motions of two prisms is overall investigated. The functional relation between the structure parameters and the exact beam scanning range is established, the capability of high-accuracy beam steering is validated together with the investigation of the scanning error sources and the nonlinear control laws, and the beam shape distortion degree under multi-parameter combinations is demonstrated. These studies can provide important references for the development of tilting double prisms.

© 2016 Optical Society of America

1. Introduction

Risley prism beam steering devices have the advantages of compact structure, small moment of inertia, good dynamic performance, large field of view (FOV) and high reliability [1–3]. The basic configuration with two identical Risley prisms rotating coaxially are widely used in free space optical communication, optical interconnection, photoelectric reconnaissance, interference measurement, and so on [4–6].

In 1960, Rosell [1] firstly proposed two Risley prisms to implement various beam scanning patterns. Then many researchers have contributed to the theoretical and applied research on Risley prisms, including inverse solutions, beam scan patterns, scan errors, blind zone, singularity problems, and Risley prism applications. For inverse solutions, Amirault and DiMarzio [7] put forward a two-step method, which simplified the variables during the inverse derivation. Tao et al. [8] established a numerical iterative inverse solution based on the damped least squares method and ensured the theoretical tracking errors less than 1 μm in both x and y directions. In 1999, Marshall’s seminal study [9] focused on the beam scan patterns through Risley prisms. By employing a first-order approximation method, Marshall simulated some interesting patterns under different wedge angles, angular rotation speeds or initial phases between prisms. Furtherly, in 2013, the exact scan patterns were presented by Schitea et al [10] with a mechanical design program, CATIA V5R20. In 2016, we [11] exhibited the blind zone resulting from the structural parameters and the singularity problems of the Risley prism system. As for the applications, researchers has developed many Risley-prism-based devices in recent years, such as confocal reflectance microscope, laser Doppler vibrometry, explosive detector, and so on [6,12,13].

In 2005, we reported the principle of tilting double-prism scanner used into free-space laser communication test [14]. The proposed scanner moves with two orthogonal tilting drive pairs [14–16], which renews the traditional Risley prism model in principle [1–4]. Because the reduction ratio from the tilting angles of the prisms to the beam deviation angle can reach a hundredfold magnitude, a common mechanical device can realize the scanning precision of submicroradian order, which is superior to the traditional rotating Risley prisms. So the tilting double-prism scanning method shows a good application potential in precision pointing fields.

In previous studies, we mainly focused on the directed function of the emergent beam, and investigated the beam scanning field, scanning precision, inverse solutions, and driving method under the hypothesis of paraxial approximation. However, some underlying technical problems should be further explored. First of all, some system parameters which haven’t been overall considered previously, may influence the beam steering performances to some extent, such as beam scanning field and exact beam scanning trajectory. And during the scanning process, the nonlinear beam scanning characteristics of the prisms has to be quantitatively described. In addition, for a typical refractive beam steering system, the beam distortion should be studied to reveal the laser energy distribution on the tracking targets. In this study, we focus on the above beam scanning problems and draw some valuable conclusions by the analysis method of multiple parameter combinations, which can provide a theoretical basis for the system design and application of tilting double-prism scanner.

2. Scanning range and scanning precision of tilting double prisms

2.1 Modelling of tilting double-prism system

A rectangular coordinate system XYZ is set as shown in Fig. 1 [17]. The beam steering system is composed of two identical circular wedge prisms with wedge angles α, refractive indexes n and the thickness of the thinnest end d0. Two prisms are sequentially named as the prism П1 and the prism П2 along the positive direction of the Z axis, respectively. At the initial state, the tilting angles of the prisms are θt1 = θt2 = 0, and the plane sides of two prisms are situated outward, which is one of the four configurations of a double-prism system proposed by Li [18]. The prism П1 can tilt around the horizontal axis L1, which is vertical to the principal cross section of prism П1 and through the center point O. Similarly, the prism П2 can tilt around the vertical axis L2, which is vertical to the principal cross section of prism П2 and through the center point O2. The distance between O and O2 is D1, and that between O2 and the screen P is D2.

 figure: Fig. 1

Fig. 1 Schematic diagram illustrating the tracking principle of tilting orthogonal double prisms.

Download Full Size | PDF

As shown in Fig. 1, At0=(0,0,1)T is the incident beam of the prism П1, At1, At2, At3 and Atf are the refracted beams after surface 11, surface 12, surface 21 and surface 22, respectively. Suppose that the tilting angles of two prisms are θt1 and θt2, the beam hits each surface of two prisms and screen P at intersection points O, Kt, Nt, Mt and Pt sequentially after refracted by two prisms. The vertical field angle ρV is defined as the angle between the projection of the emergent beam in the XOZ plane and the positive direction of Z axis, and the horizontal field angle ρH represents the angle between the projection of the emergent beam in the YOZ plane and the positive direction of Z axis.

The normal vectors of each surface of two prisms can be written as N11, N12, N21 and N22 in sequence, given by

N11=(sinθt1,0,cosθt1)T,N21=(sin(α+θt1),0,cos(α+θt1))T,N21=(0,sin(α+θt2),cos(α+θt2))T,N22=(sinθt2,0,cosθt2)T.

Then according to Snell’s law [18], each refracted beam can be expressed as Ati=(xti,yti.zti)T(i=1,2,3,f). In our previous works [15,17,19], the derivation was presented in detail, and here only the result of the emergent beam Atf is presented as follows.

Atf=(cosβt,sinβtsin(γtδ2),sinβtcos(γtδ2))T=(xtf,ytf,ztf)T,
where βt = π/2 + δ1 is the separated angle between At2 and the positive direction of the X axis, and γt = 0 is the separated angle between the projection of At2 in the YOZ plane and the positive direction of the Z axis. δ1=θt1arcsin(sini1cosαsinαn2sin2i1)α is introduced to express the deviation angle of the refractive beam At2 relative to the incident beam At0. Similarly, δ2=α+θt2arcsin(sini2cosαsinαn2¯2sin2i2)α is defined as the deviation angle of the refractive beam Atf relative to the incident beam At2 of the prism Π2. Here, n2¯=n2+(n21)cot2βt is the equivalent refractive index of prism Π2.

The vertical field angle ρV and the horizontal field angle ρH can be expressed as

ρV=arctan(xtfztf)=arctancotβtcos(γtδ2)=arctantanδ1cosδ2,
ρH=arctan(ytfztf)=γtδ2=δ2.

However, considering the structural parameters d0, D1 and D2 of the tilting double-prism system, an exact expression should be provided to describe the scanning point Pt. According to Eq. (1) and the vectors of refracted beams, the equations of each surface and each refracted beam can be solved out, and then the intersection points Kt (xtk, ytk, ztk), Nt (xtn, ytn, ztn), Mt (xtm, ytm, ztm) and Pt (xtp, ytp, ztp) can be given by

{xtk=xt1t1ytk=yt1t1ztk=zt1t1,{xtm=xt2t2+xtkytm=yt2t2+ytkztm=zt2t2+ztk,{xtn=xt3t3+xtmytn=yt3t3+ytmztn=zt3t3+ztm,{xtp=xtft4+xtnytp=ytft4+ytnztp=ztft4+ztn,
where

t1=sin(α+θt1)(dsinθt1)+cos(α+θt1)(dcosθt1)xt1sin(α+θt1)+zt1cos(α+θt1)
t2=sin(α+θt2)(ytkdsinθt2)cos(α+θt2)[ztk(D1dcosθt2)]yt2sin(α+θt2)+zt2cos(α+θt2)
t3=sinθt2ytmcosθt2(ztmD1)yt3sinθt2+zt3cosθt2
t4=D1+D2ztnztf

2.2 Scanning range of the emergent beam

Based on Eqs. (2-4) given in Section 2.1, the scanning range of the tilting double-prism system can be analyzed. In earlier researches [14, 15], the deviation angle of the emergent beam were calculated within the monotonous zone of θt1∈[0°,5°] and θt2∈[0°,8°], with no regard to other possible tilting angle ranges. According to Eq. (3) and Eq. (4), the functional relations between the vertical field angle ρV, the horizontal field angle ρH, and the tilting angles θt1, θt2 are drawn in Fig. 2, taking the situation where n = 1.517, θt1∈[−45°,45°] and θt2∈[−45°,45°] as an example. Figure 2 shows that the tilting angle θt1 mainly influences the vertical field angle ρV rather than the horizontal field angle ρH, while the tilting angle θt2 almost only influences the horizontal field angle ρH. Obviously, with the increase of the wedge angles α, the scanning range is greatly enlarged and moves away from the optical axis in both the vertical and the horizontal direction.

 figure: Fig. 2

Fig. 2 The field angle of the emergent beam. (a) the vertical field angle ρV; (b) the horizontal field angle ρH.

Download Full Size | PDF

As shown in Fig. 2, the case of α = 5° is taken as an example. Providing that θt2 is a constant, ρV increases to the maximum and then decreases along with the increase of θt1 within [−45°, 45°]. Similarly, when θt1 is a constant, ρH increases to the maximum and then decreases along with the increase of θt2 within [−45°, 45°]. Interestingly, two maxima, namely ρVmax = −2.5908° and ρHmax = −2.5925°, are both obtained at θt1 = −3.78° and θt2 = −1.17°. And the minima of ρV and ρH are both found at θt1 = θt2 = 45°, where ρVmin = −5.2406° and ρHmin = −4.5602°. So the theoretical scanning ranges are ∆ρV = ρVmaxρVmin = 2.6498°, and ∆ρH = ρHmaxρHmin = 1.9677°. In other two cases of α = 10° and 13°, the different scanning ranges with similar change trends can be obtained.

Then an experimental setup shown in Fig. 3 is developed to validate the scanning range. The parameters of the tilting double prisms are as follows: the wedge angles α = 5°, the refractive indexes n = 1.517, the clear apertures Dp = 80 mm, the thicknesses of the thinnest end d0 = 5 mm, the distance between two prisms D1 = 120 mm, and the distance between the surface 22 of prism Π2 and the receiving screen D2 = 2000 mm. The laser source with 650 nm wavelength is adjusted to be coaxial with the optical axis. The laser beam is deflected by two prisms and hits the receiving screen with two dimension grids. Then the scanning point on the grid can be measured and thus the vertical and horizontal field angle ρV and ρH can be obtained. It is noteworthy that the tilting angles are limited to θt1∈[−45°,45°] and θt2∈[−45°,45°] in the experiment.

 figure: Fig. 3

Fig. 3 Experimental setup, mainly including tilting double prism scanner, receiving screen, prism controller, laser source, laser controller and guide.

Download Full Size | PDF

Considering the structural parameters of the tilting double prism scanner, the exiting point from prism Π2 is not the center of the surface 22. However, because the distance D2 is much longer than D1, the deviation from the center of the surface 22 to the actual exiting point can be neglected. And then the vertical field ρV and the horizontal field angle ρH can be approximately calculated according to the location Pt'(x,y) of the scanning point on the receiving screen in the form of

ρV=arctan(xD2),ρH=arctan(yD2)

In the experiment, two controllable prisms are rotated to θt1 = −3.78° and θt2 = −1.17° firstly and θt1 = θt2 = 45° secondly, where the deviation angle of the beam can reach the minimum or the maximum value in theory. Then the positions of the laser beam on the receiving screen are measured. Consequently, ρVmax=2.3244° and ρHmax=2.7709° corresponding to θt1 = −3.78° and θt2 = −1.17°. And ρVmin=4.9807° and ρHmin=4.7361° corresponding to θt1 = θt2 = 45°. Hence, the measured scanning ranges are ΔρV=ρVmaxρVmin=2.6563°, and ΔρH=ρHmaxρHmin=1.9652°.Comparing the experimental results with the theoretical ones, the relative errors of ΔρV to ΔρV and ΔρH to ΔρHare less than 0.245% and 0.127% respectively, which well validate the scanning range of the tilting double prisms. However, the relatively small FOV also reveals the shortcoming of the tilting double prisms with comparison to the rotating ones.

As shown in Fig. 2, ρV and ρH are non-monotonous functions of θt1 and θt2 in the ranges of θt1∈[−45°, 45°] and θt2∈[−45°, 45°]. However, for the easy control of the tilting double prisms and to avoid multi-group inverse solutions of θt1 and θt2, the range of the tilting angles θt1∈[−45°, 45°] and θt2∈[−45°, 45°] are not appropriate. With the overall consideration of the mechanical structure, the tilting range of the prisms should be set to [0°, 45°], where functions of ρV and ρH belong to monotonic functions.

2.3 Scanning precision of the emergent beam

Based on the total deviation angle of the emergent beam, it has been theoretically confirmed that a high scanning precision can be achieved for a titling double-prism system with wedge angles α = 5° or with wedge angles α = 10° [14,15,20]. Furthermore, an experiment was conducted and strongly validated the conclusions [15]. However, considering the motion of the target, the scanning precision in the vertical and horizontal direction should be further explored. And a larger range of the tilting angle of each prism should be taken into consideration to determine an appropriate range with a relative higher scanning precision.

The scanning error of the emergent beam consists of error terms induced by the tilting error, the wedge angle error and the refraction index error of the prisms, which can be expressed as

δV=|ρVθt1|δθt1+|ρVθt2|δθt2+|ρVα|δα+|ρVn|δn,
δH=|ρHθt1|δθt1+|ρHθt2|δθt2+|ρHα|δα+|ρHn|δn,
where δV and δH are the emergent beam errors in the vertical and horizontal direction, δθt1 and δθt2 represent the tilting errors of the prisms, and δ𝛼 and δn represent the wedge angle error and the refractive index error of the prisms in sequence.

Suppose that α = 10° and n = 1.517, then the relations between the directions of the emergent beam and the tilting angles of the prisms can be drawn in Fig. 4. Figures 4(a) and 4(b) show the change laws of the partial derivatives expressed by ρVθt1 and ρHθt1 when the prism П2 is fixed. Similarly, Figs. 4(c) and 4(d) show the change laws of the partial derivatives expressed by ρVθt2 and ρHθt2 when the prism П1 is fixed.

 figure: Fig. 4

Fig. 4 The change laws of the partial derivatives between the field angles and the tilting angles. (a) and (b) represent the influences of θt1 on ρVθt1 and ρHθt1; (c) and (d) represent the influences of θt2 on ρVθt2 and ρHθt2, respectively.

Download Full Size | PDF

It can be concluded from Fig. 2 and Fig. 4 that with the increase of the tilting amplitudes of prisms in the ranges of [0, 45°], the corresponding field angles of the emergent beam are enlarged, while the absolute values of the change rate of the field angles increase, which means the scanning precision of the emergent beam reduces to a certain extent. Specially, when θt1∈[20°,45°] and θt2∈[20°,45°], the scanning precision drops sharply with the increase of θt1 and θt2. Therefore, in order to ensure a high beam steering precision, the tilting angles are limited within θt1∈[0°,10°] and θt2∈[0°,10°], where the vertical and horizontal field angles vary from −5.66° to −5.29° and −5.44° to −5.24°.

Under the above conditions, the maxima of |ρVα| and |ρHα| can both be found at θt1 = θt2 = 10°, and the values are 0.621 and 0.567, respectively. Hence, the maximal errors of ρV and ρH caused by α are 3.01 μrad and 2.75 μrad, with the hypothesis that the manufacturing precision of the wedge angle is up to 1″. Similarly, at θt1 = θt2 = 10°, |ρVn| and |ρHn| reach the maxima, of which the values are 0.196 μrad and 0.182 μrad, respectively. So the maximal errors of ρV and ρH caused by n are 1.960 μrad and 1.820 μrad, with the hypothesis that the refractive index error caused by the inhomogeneity of the optical glass is ± 1 × 10−5. Both the wedge angle error and refractive index error belong to systematic error terms, which can be corrected by the optics testing and calibration. Therefore, the random errors of the tilting mechanisms are the primary error sources that determine the scanning field angle errors, which are analyzed in detail as follows.

According to Fig. 4, |ρVθt1|, |ρVθt2|, |ρHθt1| and |ρHθt2| get their maximal values at θt1 = θt2 = 10°, and the values are 5.42 × 10−2, 3.17 × 10−4, 8.22 × 10−4 and 0.339 × 10−2, respectively. That is to say, the maximal errors of the field angles of the emergent beam are

δVδθt1×5.42%+δθt2×0.0317%,
δHδθt1×0.0822%+δθt2×0.339%,
which well demonstrates the high-accuracy scanning mechanism of the tilting double-prism system in both the vertical and the horizontal direction of the emergent beam. It is noteworthy that the above calculation on beam scanning precision are based on the condition of α = 10° and n = 1.517. Actually, a higher scanning precision can be reached if a smaller wedge angle α or a lower refractive index n is selected.

3. Analysis of beam scanning

3.1 Nonlinear problems of beam scanning

According to Fig. 2, the vertical field angle of the final emergent beam are mainly affected by θt1, and the horizontal field angle are mainly affected by θt2. Actually, a complicated nonlinear relation generally exists between the deviation angle of the emergent beam and the tilting angles of the prisms when θt1∈[0°,10°] and θt2∈[0°,10°], as mentioned in our previous work [17,19]. Suppose that one prism do not tilt, the other prism tilt individually, then the above nonlinear relations can be quantitatively described as follows. Based on the equations given in Section 2.1, Fig. 5 shows the relations between θt1, θt2 and the deviation angle ρ of the final emergent beam (including the vertical field angle ρV and the horizontal field angle ρH) with the wedge angle α = 10° and the refractive index n = 1.517.

 figure: Fig. 5

Fig. 5 The nonlinear relations between the tilting angles θt1 and θt2 of two prisms and the deviation angle ρ of the final emergent beam. (a) θt1 corresponding to ρV and ρH; (b) θt2 corresponding to ρV and ρH.

Download Full Size | PDF

As shown in Fig. 5, there are two sets of inverse tilting angle solutions for one specific combination of deviation angles ρV and ρH. But within the aforementioned chosen range of θt1∈[0°,10°] and θt2∈[0°,10°], the tilting angles θt1 and θt2 are strictly one-to-one mapped to the deviation angles ρV and ρH of the final emergent beam, so there is a unique set of the tilting angle solutions of prisms corresponding to the vertical field angle and the horizontal field angle of the final emergent beam. However, due to the nonlinear function relations between θt1 and ρV, and θt2 and ρH, the inverse solutions are preferably solved by a numerical solution method rather than the analytic ones [17, 19].

Similarly, Fig. 6 shows the nonlinear relation between the angular velocities of prisms and the change rate of deviation angle. As we can see, within the beam scanning range, the change rate of tilting angle θt1 to the deviation angle ρV nonlinearly decreases with the increase of ρV, and that of θt2 to ρH nonlinearly decreases with the increase of ρH, too. In other words, the closer the emergent beam is to the system optical axis, the faster the prisms tilt. Hence for the esay control of the prisms, during system parameter selection, an appropriate scanning range should be designed on the basis of target position estimation. In addition, the change rates of θt2 to ρV and θt1 to ρH are near zero, which in another way proves the conclusion drawn from Section 2.2 that the tilting angle θt1 mainly influences the vertical field angle ρV rather than the horizontal field angle ρH, while the tilting angle θt2 almost only influences the horizontal field angle ρH.

 figure: Fig. 6

Fig. 6 The nonlinear relation between the tilting angular velocities of prisms and the change rate of deviation angle dθ/dρ. (a) ρV corresponding to dθt1/dρV and dθt2/dρV; (b) ρH corresponding to dθt1/dρH and dθt2/dρH.

Download Full Size | PDF

Faced with the above nonlinear problems, some appropriate strategies must be taken for the real-time nonlinear control. Last year, we proposed a nonlinear control scheme to solve this problem [19]. A tilting double-prism scanner driven by cam-based mechanism was designed to transform the complicated control procedure of the drive motor to the profile curve design of the cam. Compared to the reciprocating rotational motor movement in earlier rotating or tilting Risley prism scanners [5–7,14–16], this cam-based continuous rotational motor driving method is easier to be performed, and could achieve a sufficiently high oscillatory frequency. This method was confirmed to be feasible for a steady scanning trajectory.

3.2 Beam scanning trajectory

Suppose that the incident beam vector is At0=(0,0,1)T, according to Eqs. (5-9), the various scanning trajectories can be drawn when the prisms tilt at different angular velocity combinations within the range of 0°~10° (In the following part of the paper, the ranges of the tilting angles are set to 0°~10°without special notes).

In Fig. 7(a), the beam scanning trajectory curve is viewed when the tilting angular velocities of the prisms keep uniform and ωt2 = 2ωt1. The tilting angles of the prisms are θt1=|t10|+10 and θt2=|2t10|+10, respectively. As we can see, the trajectory is a closed curve, and the scanning point moves along the curve circularly. Similarly, Figs. 7(b) and 7(c) show the beam scanning trajectories with ωt2 = 4ωt1, and the tilting angles of the prisms are θt1=|t10|+10 and θt2=|4t10|+10, respectively. During the scanning process, the scanning point firstly moves along the path shown in Fig. 7(b), basically toward the positive direction of the X axis, and then returns to the starting point along the identical path shown in Fig. 7(c). Besides the above uniform tilting angular velocities, we simulate several scanning curves with non-uniform velocities according to the sine or cosine function laws. In Figs. 7(d) and 7(e), the tilting angles θt1 are the same, and θt1=5sin(πt/5)+5, while θt2=|2t10|+10 and θt2=5cos(πt/5)+5, respectively. The scanning trajectories in Figs. 7(d) and 7(e) are two closed curves, and the scanning points move along the curve circularly. It is noteworthy that the former has sharp points on the trajectory, while the latter is a relatively smooth curve over a full cycle due to sine function for ωt1 and cosine function for ωt2. In general, the inflection points on these scanning trajectories as shown in Figs. 7(a)-7(d) reflect the discontinuous differentiability of the tilting angle function, and the closed trajectory curves indicate the beam scanning periodicity caused by the different tilting cycles within the limited tilting angle range of 0°~10°.

 figure: Fig. 7

Fig. 7 Beam scanning trajectories with different tilting angular velocities, where the red arrows show the moving direction of the scanning point. These tilting angular velocities are chosen arbitrarily, while in practice the tilting angles and velocities should be determined based on specific applications. (a) uniform ωt1, uniform ωt2 and ωt2 = 2ωt1; (b) and (c) uniform ωt1, uniform ωt2 and ωt2 = 4ωt1; (d) sine function for ωt1 and uniform ωt2; (e) sine function for ωt1 and cosine function for ωt2.

Download Full Size | PDF

It can be seen from Fig. 7 that two prisms with different angular velocity combinations can produce various trajectory curves. In fact, according to the given target trajectory, the corresponding tilting angle curves of prisms can be obtained using an inverse solution method [19]. Through the precise nonlinear control or the mechanism transmission design, any desired beam scanning trajectories within the scanning field can be generated in theory.

3.3 Influences of system structure parameters on scanning field

In the design of the tilting double-prism system, besides the scanning precision, the position and size of the scanning field on the receiving device are also significant indexes. In Section 2.2, we discuss the scanning range of the emergent beam to illustrate the vertical field angle ρV and the horizontal field angle ρH. However, with the change of some structure parameters, such as D1, d0 and D2, the scanning point position on the receiving device can make different shifts, while the emergent beam direction has no changes. Hence in this section, we will discuss the influences of D1, d0 and D2 on the scanning field with the wedge angles α = 10°, the refractive indexes n = 1.517, and the clear apertures Dp = 80 mm, respectively.

Figure 8 shows the scanning field with structure parameters of D1 = 100 mm, D2 = 400 mm and d0 = 5 mm. The real scanning field is a parallelogram-like area with the angles between each adjacent edges very close to a right angle. For convenience, the biggest rectangle area in the real scanning field is taken as the scanning field. In this example, the scanning field locates at the area of xtp∈[−47.17, −44.82] and ytp∈[−39.63, −37.44], the scanning area A = 2.35 mm × 2.19 mm = 5.15 mm2 and the center point coordinate is (−45.99, −38.53) mm.

 figure: Fig. 8

Fig. 8 The scanning field of a tilting double-prism scanner. The area enclosed by the red rectangle, which is the largest rectangle within the parallelogram-like area, is taken as the scanning field of this scanner.

Download Full Size | PDF

In Table 1, the scanning field parameters are viewed under the different separation distance D1 between two prisms when D2 = 400 mm and d0 = 5 mm. Suppose that D2 is constant, the change of D1 only affects the coordinate range of the scanning field in the X direction, while that in the Y direction remains unchanged, as shown in Table 1. The reason is that the deviation of the final emergent beam in Y direction is mainly caused by the tilt of the prism П2, and the optical path after surface 21 has nothing to do with the distance between two prisms D1. With the increase of D1, the scanning range in the X direction increases, the area of the scanning field increases and the center point of the scanning field moves along the negative direction of the X axis.

Tables Icon

Table 1. Parameters of Scanning Field under Different D1 (mm)

Similarly, Table 2 shows the parameters of scanning field which are obtained under different d0 when D1 = 100 mm and D2 = 400 mm. When D1 and D2 keep invariant, with the increase of d0, the coordinate range of the scanning field in the X direction decreases, while that in the Y direction increases. Correspondingly, the X coordinate value of the center point of the scanning field increases but the Y coordinate value decreases.

Tables Icon

Table 2. Parameters of Scanning Field under Different d0 (mm)

In addition, if D1 and d0 keep invariant while D2 changes, it is easy to conclude that the farther the receiving screen is, the larger the scanning area becomes, and the farther the center point of the scanning field deviates from the origin of the screen.

4. Beam distortion analysis

In general, for a refractive optical system, the beam distortion should be taken into consideration in practical applications. In previous researches, the beam distortion of rotation Risley prisms was studied using the vector refraction theorem [21, 22]. In 2016, we further presented a detailed analysis about it considering the effects of system parameters and prism configurations [23]. The above works shows that the distortion effect on the beam shape is relatively weak in early double-prism systems with shallow wedge angles and small refractive indexes. But with the larger wedge angles and higher refractive indexes applied in the prism scanners, the distortion degree become too severe to be neglected anymore. However, as for a tilting double-prism system, the beam distortion is also an important issue in some application situations, and a comprehensive analysis should be performed to reveal the beam distortion laws.

In this section, a ray-tracing method is used to analyze the beam shape distortion. The incident beam At0 parallel to the Z axis is prescribed to be circular with radius r = 20 mm, and the central axis of the incident beam intersects with the surface 11 at the point (10, 0, 0). Some necessary system parameters defined in Section 2.1 are reset as below: D1 = 400 mm, D2 = 400 mm, n = 3, α = 10°, d0 = 10 mm and Dp = 400 mm. Therefore, the beam edge on the prism surface 11 can be written as (unit is millimeters)

{x=20cos(θ)y=10+20sin(θ)z=0,
where θ∈(0°, 360°).

By utilizing the ray-tracing method and the vector refraction theorem, the intersection lines between the beam edge and each prism surfaces can be expressed by coordinate equations sequentially. This process is very similar to the forward derivation of the beam propagation [15]. Here the beam distortion degree ε is defined as the ratio of the maximal distortion value of the emergent beam in radial direction to the initial diameter of the incident beam.

Figure 9 indicates that the distortion degree ε reaches 8.46%, 9.52%, 13.01% and 20.30% respectively under the different tilting angle combinations of two prisms, namely θt1 = θt2 = 0°, θt1 = 0° and θt2 = 5°, θt1 = 0° and θt2 = 10°, θt1 = 10° and θt2 = 10°. Obviously, the beam exiting from the tilting double prisms is squeezed in one direction but stretched in the mutual perpendicular direction, which is similar to the results in rotation Risley prisms [23]. It is also evident that the distortion degree ε of the emergent beam depends on the tilting angles of two prisms.

 figure: Fig. 9

Fig. 9 The beam shape distortion induced by double prisms with different tilting angle combinations. Red curves with “°” signs and blue curves with “×” signs represent the original beam shapes and the distorted ones, respectively. (a) θt1 = 0°, θt2 = 0°; (b) θt1 = 0°, θt2 = 5°; (c) θt1 = 0°, θt2 = 10°; (d) θt1 = 10°, θt2 = 10°.

Download Full Size | PDF

Furthermore, when the wedge angle α, the refractive index n, the thickness of the thinnest end d0 and the distance D1 change in sequence, the distortion laws of the beam can be observed in comparison with the initial circular shape. Providing the tilting angles of prisms are θt1 = 0° and θt2 = 0°, θt1 = 0° and θt2 = 5°, θt1 = 0° and θt2 = 10°, θt1 = 10° and θt2 = 10° respectively, the corresponding beam distortion degree ε under different α, n, d0 and D1 can be calculated, and then the distortion laws are clearly revealed in Fig. 10.

 figure: Fig. 10

Fig. 10 The influence of system parameters on beam distortion. (a) refractive index n; (b) wedge angle α; (c) thickness of the thinnest end d0; (d) distance D1 between O and O2.

Download Full Size | PDF

Seen from Fig. 10, we can draw some valuable conclusions. Both the higher refractive index and the larger wedge angle can aggravate the distortion degree on beam shape, while the thicknesses of thinnest end and the distance between double prisms are verified to be nearly irrelevant factor. And in general, the distortion degree increases with the tilting angles. In the following comparisons, θt1 = 0° and θt2 = 5°. Let α = 15°, d0 = 10 mm and D1 = 400 mm, when n increases from 2 to 3, ε correspondingly increase from 4.42% to 7.63%. Similarly, let n = 3, d0 = 10 mm and D1 = 400 mm, ε equals to 3.18%, 7.63% and 23.04%, in response to the incremental α values of 5°, 10° and 15°. As for the distance D1 between two prisms, the distortion degree ε keeps uniform no matter how D1 varies. Moreover, the influence of d0 on the distortion degree ε can be neglected. Generally, there is a contradiction between wider FOV with higher or larger α and less beam distortion with lower n or smaller α, and a balance between these requirements has to be comprehensively considered.

5. Conclusions

A pair of tilting double prisms, as one of the most promising scanning methods, can realize much higher beam-steering accuracy than the traditional rotation Risley-prism scanning systems. The beam scanning range and scanning precision, as well as the beam distortion, are the basic beam performance indexes which condition the use fields of this scanning system. In the paper, we discuss the influences of key system parameters on the beam scanning properties, including the wedge angle, the thickness and the refractive index of the prism, the distance between two prisms, and the distance between the scanner and the receiving device. As for the specific scanning field on the receiving screen, its size and position are determined by the change of the above parameters. In order to regularize the motion control, the nonlinear relation between the tilting angles of the prisms and the deviation angle of the emergent beam is thoroughly investigated, and some possible solutions are proposed. Moreover, the beam distortion is irrelevant to the distance between two prisms, while is greatly influenced by the wedge angle and the refractive index. These results can provide the foundation for the design and application of the tilting double-prism scanning system.

Funding

National Natural Science Foundation of China (NSFC) (51375347, 61675155).

References and links

1. F. A. Rosell, “Prism scanner,” J. Opt. Soc. Am. 50(6), 521–526 (1960). [CrossRef]  

2. Y. Yang, “Analytic solution of free space optical beam steering using Risley prisms,” J. Lightwave Technol. 26(21), 3576–3583 (2008). [CrossRef]  

3. G. F. Marshall, Handbook of Optical and Laser Scanning, 2nd ed. (CRC Press, 2011).

4. C. R. Schwarze, R. Vaillancourt, D. Carlson, E. Schundler, T. Evans, and J. R. Engeletc, “Risley-prism based compact laser beam steering for IRCM, laser communications, and laser radar,” http://www.optra.com/images/TP-Compact_Beam_Steering.pdf.

5. V. F. Duma, J. P. Rolland, and A. G. Podoleanu, “Perspectives of optical scanning in OCT,” Proc. SPIE 7556, 75560B (2010). [CrossRef]  

6. S. J. Rothberg and M. Tirabassi, “Development of a scanning head for laser Doppler vibrometry (LDV) using dual optical wedges,” Rev. Sci. Instrum. 84(12), 121704 (2013). [CrossRef]   [PubMed]  

7. C. T. Amirault and C. A. Dimarzio, “Precision pointing using a dual-wedge scanner,” Appl. Opt. 24(9), 1302–1308 (1985). [CrossRef]   [PubMed]  

8. X. Tao, H. Cho, and F. Janabi-Sharifi, “Active optical system for variable view imaging of micro objects with emphasis on kinematic analysis,” Appl. Opt. 47(22), 4121–4132 (2008). [CrossRef]   [PubMed]  

9. G. F. Marshall, “Risley prism scan patterns,” Proc. SPIE 3787, 74–86 (1999). [CrossRef]  

10. A. Schitea, M. Tuef, V. F. Duma, and A. M. Vlaicu, “Modeling of Risley prisms devices for exact scan patterns,” Proc. SPIE 8789, 878912 (2013). [CrossRef]  

11. A. Li, W. Sun, W. Yi, and Q. Zuo, “Investigation of beam steering performances in rotation Risley-prism scanner,” Opt. Express 24(12), 12840–12850 (2016). [CrossRef]   [PubMed]  

12. W. C. Warger 2nd and C. A. DiMarzio, “Dual-wedge scanning confocal reflectance microscope,” Opt. Lett. 32(15), 2140–2142 (2007). [CrossRef]   [PubMed]  

13. E. Schundler, D. Carlson, R. Vaillancourt, J. R. Dupuis, and C. Schwarze, “Compact, wide field DRS explosive detector,” Proc. SPIE 8018, 80181O (2011). [CrossRef]  

14. A. Li, J. Sun, L. Wang, and L. Liu, “Submicroradian accuracy scanning system with a double-wedge rotating around the orthogonal axes,” Proc. SPIE 5892, 58921M (2005). [CrossRef]  

15. A. Li, L. Liu, J. Sun, X. Zhong, L. Wang, D. Liu, and Z. Luan, “Research on a scanner for tilting orthogonal double prisms,” Appl. Opt. 45(31), 8063–8069 (2006). [CrossRef]   [PubMed]  

16. L. Liu, L. Wang, J. Sun, Y. Zhou, X. Zhong, Z. Luan, D. Liu, A. Yan, and N. Xu, “An Integrated Test-Bed for PAT Testing and Verification of Inter-Satellite Lasercom Terminals,” Proc. SPIE 6709, 670904 (2007). [CrossRef]  

17. A. Li, Y. Ding, Y. Bian, and L. Liu, “Inverse solutions for tilting orthogonal double prisms,” Appl. Opt. 53(17), 3712–3722 (2014). [CrossRef]   [PubMed]  

18. Y. Li, “Closed form analytical inverse solutions for Risley-prism-based beam steering systems in different configurations,” Appl. Opt. 50(22), 4302–4309 (2011). [CrossRef]   [PubMed]  

19. A. Li, W. Yi, W. Sun, and L. Liu, “Tilting double-prism scanner driven by cam-based mechanism,” Appl. Opt. 54(18), 5788–5796 (2015). [CrossRef]   [PubMed]  

20. A. Li, X. Jiang, J. Sun, L. Wang, Z. Li, and L. Liu, “Laser coarse-fine coupling scanning method by steering double prisms,” Appl. Opt. 51(3), 356–364 (2012). [CrossRef]   [PubMed]  

21. W. Mao and Y. Xu, “Distortion of optical wedges with a large angle of incidence in a collimated beam,” Opt. Eng. 38(4), 580–585 (1999). [CrossRef]  

22. J. Sun, L. Liu, M. Yun, L. Wan, and M. Zhang, “Distortion of beam shape by a rotating double prism wide-angle laser beam scanner,” Opt. Eng. 45(4), 043004 (2006). [CrossRef]  

23. A. Li, Q. Zuo, W. Sun, and W. Yi, “Beam distortion of rotation double prisms with an arbitrary incident angle,” Appl. Opt. 55(19), 5164–5171 (2016). [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 Schematic diagram illustrating the tracking principle of tilting orthogonal double prisms.
Fig. 2
Fig. 2 The field angle of the emergent beam. (a) the vertical field angle ρV; (b) the horizontal field angle ρH.
Fig. 3
Fig. 3 Experimental setup, mainly including tilting double prism scanner, receiving screen, prism controller, laser source, laser controller and guide.
Fig. 4
Fig. 4 The change laws of the partial derivatives between the field angles and the tilting angles. (a) and (b) represent the influences of θt1 on ρ V θ t1 and ρ H θ t1 ; (c) and (d) represent the influences of θt2 on ρ V θ t2 and ρ H θ t2 , respectively.
Fig. 5
Fig. 5 The nonlinear relations between the tilting angles θt1 and θt2 of two prisms and the deviation angle ρ of the final emergent beam. (a) θt1 corresponding to ρV and ρH; (b) θt2 corresponding to ρV and ρH.
Fig. 6
Fig. 6 The nonlinear relation between the tilting angular velocities of prisms and the change rate of deviation angle dθ/dρ. (a) ρV corresponding to dθt1/dρV and dθt2/dρV; (b) ρH corresponding to dθt1/dρH and dθt2/dρH.
Fig. 7
Fig. 7 Beam scanning trajectories with different tilting angular velocities, where the red arrows show the moving direction of the scanning point. These tilting angular velocities are chosen arbitrarily, while in practice the tilting angles and velocities should be determined based on specific applications. (a) uniform ωt1, uniform ωt2 and ωt2 = 2ωt1; (b) and (c) uniform ωt1, uniform ωt2 and ωt2 = 4ωt1; (d) sine function for ωt1 and uniform ωt2; (e) sine function for ωt1 and cosine function for ωt2.
Fig. 8
Fig. 8 The scanning field of a tilting double-prism scanner. The area enclosed by the red rectangle, which is the largest rectangle within the parallelogram-like area, is taken as the scanning field of this scanner.
Fig. 9
Fig. 9 The beam shape distortion induced by double prisms with different tilting angle combinations. Red curves with “°” signs and blue curves with “×” signs represent the original beam shapes and the distorted ones, respectively. (a) θt1 = 0°, θt2 = 0°; (b) θt1 = 0°, θt2 = 5°; (c) θt1 = 0°, θt2 = 10°; (d) θt1 = 10°, θt2 = 10°.
Fig. 10
Fig. 10 The influence of system parameters on beam distortion. (a) refractive index n; (b) wedge angle α; (c) thickness of the thinnest end d0; (d) distance D1 between O and O2.

Tables (2)

Tables Icon

Table 1 Parameters of Scanning Field under Different D1 (mm)

Tables Icon

Table 2 Parameters of Scanning Field under Different d0 (mm)

Equations (15)

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

N 11 = ( sin θ t1 ,0,cos θ t1 ) T , N 21 = ( sin( α+ θ t1 ),0,cos( α+ θ t1 ) ) T , N 21 = ( 0,sin( α+ θ t2 ),cos( α+ θ t2 ) ) T , N 22 = ( sin θ t2 ,0,cos θ t2 ) T .
A tf = ( cos β t ,sin β t sin( γ t δ 2 ),sin β t cos( γ t δ 2 ) ) T = ( x tf , y tf , z tf ) T ,
ρ V =arctan( x tf z tf )=arctan cot β t cos( γ t δ 2 ) =arctan tan δ 1 cos δ 2 ,
ρ H =arctan( y tf z tf )= γ t δ 2 = δ 2 .
{ x tk = x t1 t 1 y tk = y t1 t 1 z tk = z t1 t 1 ,{ x tm = x t2 t 2 + x tk y tm = y t2 t 2 + y tk z tm = z t2 t 2 + z tk ,{ x tn = x t3 t 3 + x tm y tn = y t3 t 3 + y tm z tn = z t3 t 3 + z tm ,{ x tp = x tf t 4 + x tn y tp = y tf t 4 + y tn z tp = z tf t 4 + z tn ,
t 1 = sin( α+ θ t1 )( dsin θ t1 )+cos( α+ θ t1 )( dcos θ t1 ) x t1 sin( α+ θ t1 )+ z t1 cos( α+ θ t1 )
t 2 = sin( α+ θ t2 )( y tk dsin θ t2 )cos( α+ θ t2 )[ z tk ( D 1 dcos θ t2 ) ] y t2 sin( α+ θ t2 )+ z t2 cos( α+ θ t2 )
t 3 = sin θ t2 y tm cos θ t2 ( z tm D 1 ) y t3 sin θ t2 + z t3 cos θ t2
t 4 = D 1 + D 2 z tn z tf
ρ V =arctan( x D 2 ), ρ H =arctan( y D 2 )
δ V =| ρ V θ t1 | δ θ t1 +| ρ V θ t2 | δ θ t2 +| ρ V α | δ α +| ρ V n | δ n ,
δ H =| ρ H θ t1 | δ θ t1 +| ρ H θ t2 | δ θ t2 +| ρ H α | δ α +| ρ H n | δ n ,
δ V δ θ t1 ×5.42%+ δ θ t2 ×0.0317%,
δ H δ θ t1 ×0.0822%+ δ θ t2 ×0.339%,
{ x=20cos(θ) y=10+20sin(θ) z=0 ,
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.