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

Theoretical and experimental research on temperature-induced surface distortion of deformable mirror

Open Access Open Access

Abstract

In a well-manufactured deformable mirror (DM), the temperature-induced distortion (TID) has been reported exist on the surface shape of the DM, when the working environment temperature is not equal to that of the manufacturing environment. The DM could not effectively correct this actuator-corresponding TID and the correction ability of the DM would be limited. In this paper, the the TID’s essential mechanism is analyzed systematically based on the thermal stress characteristics. An efficient method based on an auxiliary temperature compensation module (TCM) and a hybrid closed-loop control algorithm are presented accordingly. A finite element model is built to evaluate the TID characteristics and the compensation capability of the TCM. In the simulation, by using the TCM, the the DM’s improved surface shape does not contain the dynamic high-frequency distortion caused by the actuators tilt. In the experiment, which uses a designed TCM and a hybrid closed-loop control algorithm, a DM’s TID is effectively depressed and a well-compensated DM surface shape is finally achieved.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

At present, adaptive optics (AO) has been the key technology in Inertial Confinement Fusion (ICF) [1], high power laser [2,3], high resolution imaging [4,5] and atmosphere optics [6]. The deformable mirror (DM) is used to correct the wavefront distortion in an AO system. Its correction ability is influenced by the stability of the surface shape under different environment and the local high frequency distortion (HFD). The adverse impacts of the high frequency distortion in fitting residual have been reported. For example, the far field pattern is far from smooth with plenty of discrete minor spots caused by the uncorrected high-order aberrations in a high-power laser system such as the ShenGuang-III Facility (SG-III) and the National Ignition Facility (NIF), which are used for the inertial confinement fusion in China and the United States, respectively [7]. Many manufacture techniques have been adopted to restrain the static and dynamic HFD in the production stage [8,9]. For example, the dynamic HFD is eliminated by keeping the actuators strictly perpendicular to the steel base [8]. The adhesive curing induced static HFD is eliminated by inserting the transition layer between the mirror and the actuators in the DM [9].

However, the surface distortion could still occur in a well-manufactured DM when the temperature of the DM in the working environment (Te) is not equal to the temperature in the manufacturing environment (Tm). The TID of the DM has been investigated by numerical analyses and experiments [10–14]. For the bimorph DM, the temperature gradient appearing in the mirror due to low thermal conductivity of the mirror results in the surface shape distortion. The distortion could be depressed by applying a Safronov’s cooling system, and the depress effect could be improved by inserting a middle metal coating conduction layer [10]. Investigation shows that the thermal induced deformation of a well-manufactured screen-printed unimorph DM varies with the load method. A homogeneous load induces the focusing of the mirror while an inhomogeneous load will induce the defocusing of the mirror [11]. The deformation induced by the homogeneous load is influenced by the mounting materials, while the deformation induced by the inhomogeneous load depends on the copper-layer thickness. For the membrane DM, the surface shape of the mirror could be affected by the convection when the membrane temperature is above 35 [12]. To withstand 10 kW laser power and maintain the thermal characterization, a configuration of a unimorph DM is developed by mounting a cooling cavity on the back side of the DM [13].

The TID existing on the surface of the DM is one type of the actuator-corresponding dynamic HFD [14,15]. As reported, when the ambient temperature rises/falls 2, the PV value of the surface shape of a well-mounted DM would increase 1.4 μm [15], which could not be acceptable in a high-power laser system (e.g. the SG-III and the NIF). Restricted by the structure and the mechanism, the DM with limited stacked array actuators is only used to correct the low order wavefront aberrations and has limitations to correct the TID effectively [8]. Thus, in the application of an AO system, the TID will remain in the residual surface shape of the DM after correction, which obviously limits the correction ability of the DM. The uncorrectable TID will cause the degeneration of the beam quality of the optics system and the energy dispersion of the far-field spot. Moreover, under certain condition, the TID could even result in the unexpected shift of the focal point when the laser is in operation, which is very dangerous for the optical elements [9]. Thus, the application field of a DM with TID is strongly restricted by the working temperature and could not be extended to the working environment with a wider ambient temperature range.

Many techniques have been proposed to reduce the TID in the structural design, such as increasing the thickness of the base and the mirror, increasing the actuator numbers and choosing the materials of base and mirror with small thermal expansion coefficient difference [14]. The goal of all the techniques is to optimize the structure parameters under the design stage to weaken the TID, as the essential reason of the TID of the NIF-type DM hasn’t been investigated systematically. In fact, if Te is not equal to Tm, the TID always exists in the DM even the PV value of the TID is depressed by the optimization skills mentioned above. Therefore, proposing a disjunctive TID compensation method, which could be applicable for the manufactured DM, is of vital.

In this paper, the essential reason of the TID of the NIF-type DM is investigated systematically. Based on the analysis result, an efficient method based on an auxiliary TCM and a hybrid close-loop control algorithm is presented accordingly. In section 2, a simplified structure model of the DM is established. The mechanism of the TID is analyzed through the derivation of the residual thermal stress according to the simplified model. Simulation results show that the tilt of the actuators caused by the thermal stress difference of the mirror and the steel base is the essential reason of the TID. Finally, according to the TID’s mechanism, the mathematical equations of a disjunctive TCM is acquired for proving the feasibility to compensate the TID. In section 3, a finite element model is built to investigate the influence of various parameters on the TID, including the ambient temperature, the thermal expansion coefficient difference between the mirror and steel base, and the TCM temperature. The simulation results indicate the TID could be effectively eliminated by the TCM. In section 4, the experiment results show that the TID is well depressed and almost vanishes in the self-compensation residues with the TCM working.

2. NIF-type DM structure and TID mechanism

Various DMs have been proposed to meet different application requirements [16–22], such as the liquid crystal spatial light modulator [19,20], the MEMS DM [21,22] and the traditional NIF-type DM [7]. Due to the advantages of large aperture, wide dynamic range and high damage threshold, the NIF-type DM is widely used in high power laser system and ICF system.

A NIF-type DM is manufactured to analyze the mechanism of the TID, and the simplified structure is shown Fig. 1. The DM, which is a typical discrete PZT driven, continuous mirror DM, consists of an 84mm × 84mm × 1mm plate BK7 mirror, 49 stacked PZT actuators and an 84mm × 84mm × 40mm stainless-steel base. The diameter and height of actuators are 5mm and 36mm, respectively. The 49 actuators are square array distributed with 12mm spacing. The optical mirror is bonded to the PZT actuators, which are rigidly connected to the steel base. The thermal expansion coefficients of the mirror and the steel base are 7.1 × 10−6/K and 17.2 × 10−6/K, respectively. The stress of the mirror and the steel base varying with the change of the working environment temperature is different. As a result, the residual thermal stress at the joints between the actuators and the mirror might lead to the unwanted TID.

 figure: Fig. 1

Fig. 1 (a) Simplified schematic of the NIF-type deformable mirror structure. (b) The XOY coordinates and the distribution of the actuators.

Download Full Size | PDF

Two coordinate systems shown in the Fig. 1(b) and Fig. 2 are established to analyze the TID. Considering the symmetry of the structure of the DM [Fig. 1], the local surface of the actuator-mirror contact surface in the first octant could be used to represent other three octants when Te is not equal to Tm. An actuator in the first octant, the (i, j)th actuator, is selected as the analysis object. Where i represents the row number of the selected actuator along the X-direction, and j represents the column number of the selected actuator along the Y-direction, while the number of the central actuator is set (0,0). Point Ai,j is the center of the actuator-mirror contact surface, and point Bi,j is the center of the actuator-base contact surface. (xi,yj,0) and (xi,yj,h) are the coordinates of point Ai,j and point Bi,j, respectively, where xi=ia,yj=ja ,and h is the height of the actuator in the manufacturing environment. The dash line Ai,jBi,j¯ represents the original position of the (i, j)th actuator at Tm. When Te is not equal to Tm, due to the thermal expansion and contraction characterization, point Ai,j moves to point Ai,j' and point Bi,j moves to point Bi,j'. The solid line Ai,j'Bi,j'¯ represents the position of the (i, j)th actuator at Te.

 figure: Fig. 2

Fig. 2 (a) The XYZ coordinates and the total displacements of the actuator. (b) The XOY and X'O'Y' coordinates and the shear displacement of the mirror. (c) The Y'O'Z' section and the normal displacement of the mirror.

Download Full Size | PDF

By applying the notation of linear thermal expansion coefficient, the coordinates differences between point Ai,j and point Ai,j' are written as Eqs. (1) and (2), while Eqs. (3) and (4) express the coordinates differences between point Bi,j and point Bi,j' .

δiA=xiA'xi=ΔTeiaαG
δjA=yjA'yj=ΔTejaαG
δiB=xiB'xi=ΔTeiaαB
δjB=yjB'yj=ΔTejaαB

Here δiA is the temperature induced deformation value along the X-direction of point Ai,j. δjA is the temperature induced deformation value along the Y-direction of point Ai,j. δiB and δjB is the corresponding value of point Bi,j. αG and αB are the linear thermal expansion coefficients of the mirror and steel base, respectively. ΔTe represents the difference between Te and Tm, which is expressed as Eq. (5).

ΔTe=TeTm

Considering the practical operation environment, the temperature of the DM in the working environment Te is hardly equal to the temperature in the manufacturing environment Tm. When the temperature Te of the DM is not equal to the original manufacturing temperature Tm, ΔTe0,  δiAδiB and δjAδjB, which indicate that solid line Ai,j'Bi,j'¯ is not normal to the mirror. Consequently, the actuator is tilted and stretched. The tensile stress appears in the actuator to resist the tensile deformation, which is influenced by two factors: one is the reaction force when the actuator is stretched, another is the heat expansion and contraction. The contraction along the ideal tilted axis Ai,j'Bi,j'¯ induced by the tensile stress is supposed as the total reverse displacements Δhij. As shown in Figs. 2(b) and 2(c), the second coordinate system X'O'Y' is defined by selecting point Ai,j as the origin O' and selecting the direction of the vector from point Ai,j to point Ai,j' as the forward direction of the X' axis. Figure 2(c) shows the cross section Y'O'Z' and the normal displacements Δhijz (blue line) caused by Δhij. The shear displacements Δhijx and Δhijy are illustrated by the blue line in Fig. 2(b). The normal displacements Δhijz, the shear displacements Δhijx and Δhijy could be expressed in Eq. (6).

{Δhijz=Δhijcos(θi,j)Δhijx=Δhijsin(θi,j)sin(αi,j)Δhijy=Δhijsin(θi,j)cos(αi,j)

Where the angle αi,j is associated with the row number i and the column number j [Fig. 2(b), Eq. (7)]. αi,j is the incline angle between the Z-direction and the direction of the tilted axis Ai,j'Bi,j'¯ [Fig. 2(c), Eq. (8)].

tan(αi,j)=δiAδjA=ij(j0)
tan(θi,j)=δBi,jδAi,jh
δAi,j=δiA2+δjA2

Specially, αi,j is equal to π/2 when j=0. δAi,j is the total displacement from point Ai,j to point Ai,j' on the XOY plane [Eq. (9)], and all the subscript A could be replaced by B in Eq. (9) to represent the total displacement from point Bi,j to point Bi,j'.

For a DM, the surface shape of the mirror is driven and controlled by the vertical movements of the PZT actuators to generate the conjugate surface of the distorted wavefront. Therefore, the normal displacement Δhijz [Eq. (6)] appearing on the mirror surface could be corrected by the pushing and pulling of the actuators vertically, while the shear displacements Δhijx and  Δhijy of the actuators could not be compensated and would result in the dynamic HFD in the residue [8]. Therefore, in a DM, the TID in the correction residual could be eliminated if Δhijx and  Δhijy are always equal to zero even if the ΔTe is not equal to zero [Eq. (6)].

In order to keep Δhijx and Δhijy always equal to zero, the angle θi,j in Eq. (6) should be set to zero. By introducing Eqs. (1)–(4) and (9), Eq. (8) could be rewritten as Eq. (10).

tan(θi,j)=ai2+j2(ΔTeBαBΔTeGαG)h

Where the structure parameters α is not equal to zero. ΔTeB is the difference between the DM base’s temperature in the working environment and Tm, while ΔTeG is the difference between the DM mirror’s temperature in the working environment and Tm. Obviously, in the working environment with steady temperature field, ΔTeB and ΔTeG are both equal to ΔTe. As the materials of the steel base and the mirror in a DM could not be the same, the thermal expansion coefficient αG is not equal to αB. From Eq. (10), in the working environment with the temperature not equal to the original manufacture temperature (most probably happening in the real practice), the angle θi,j could still be kept to zero if the relative parameters are set to satisfy Eq. (11).

ΔTeB=αGαBΔTeG

A TCM is designed to regulate the temperature field of the steel base to achieve the target ΔTeB, as shown in Fig. 3, which could be installed to the rear surface of the steel base for a manufactured DM. The steel base temperature could be regulated by the heat conduction with the TCM. The mirror and the steel base are indirectly connected by low thermal conductivity PZT-actuators, and the area of the surface contacting with air is much larger than the area of mirror actuators contacting surface. According to the theory of heat conduction, the mirror temperature mainly depends on the convective heat transfer with air, which almost equals to the temperature of the DM in the working environment, even the temperature of the steel base is regulated by the auxiliary TCM. Therefore, Eq. (11) could be rewritten as Eq. (12).

 figure: Fig. 3

Fig. 3 (a)The DM with the TCM. (b) Distribution of four TECs.

Download Full Size | PDF

ΔTeB=αGαBΔTe

To improve the heat conductivity, the heat-conducting silicon grease is applied on the contacting layer between the steel base and the TCM. The designed TCM consists of four-square array distributed TECs (single TEC size is 40mm × 40mm × 4mm, and the combined size is 80mm × 80mm × 4mm, 12V/5A, TEC1204) [Fig. 3(b)], a water-cooling block, and a temperature control circuit. The stable temperature of the steel base depends on Te and the TCM temperature Ttcm. Thus, the stable ΔTeB could be expressed as Eq. (13) based on the theory of heat conduction.

ΔTeB=f(Te,TTCM,Tm)

Apparently, for the stable working environment with the constant value of Te, the parameter ΔTeB could always be controlled solely and directly by adjusting the temperature of the TCM. Based on this method, the angle θi,j in Eq. (6) could be kept zero and the TID in the residual could be eliminated. Thus, the application field of a DM would be extended to an unrestricted working environment with a wider ambient temperature range.

3. Simulation analysis of temperature induced distortion

To evaluate the TCM’s compensation capability on the TID of the DM, a finite element model is built in COMSOL Multiphysics software [23,24]. In the simulation model, the DM consists of a continuous plate BK7 mirror (84mm × 84mm × 1mm full size), a stainless-steel base (84mm × 84mm × 1mm full size) and 49 square distributed discrete PZT posts (5mm-diameter, 36mm-length). Primary parameters of the materials as constants in small temperature variation range are listed in Table 1.

Tables Icon

Table 1. Material parameters in the finite element simulation

The whole model is built using the thermal stress interface from the structural mechanics module. In the simulation, the degrees of freedom for the four corners of the steel base are set to zero, and the Z-direction of the steel base’s rear surface is also set to zero. The initial surface shape of the mirror is set to be an ideal plane without any distortion. The heat transfer is by means of the natural convective heat flux. For the heating boundary condition, the convective heat transfer coefficient of all outer surfaces is set 10  W/(m2K), while the coefficient of the base TCM contacting surface is set 25  W/(m2K). The original manufacturing environment temperature Tm of the DM is set 20 .

3.1 TID characteristics of DM

In the simulation, the working temperature of the DM is set ΔTedeviation from the original manufacturing environment temperature Tm. Figures 4(a) and 4(b) show the temperature gradient map and the displacement map of the whole DM without any distortion compensation. From Fig. 4(a), when the temperature of the steal base is not controlled, ΔTeB and ΔTeG are both equal to 5 (ΔTe). The thermal expansion coefficients difference between the mirror and the steal base will result in the expansion mismatch and consequently lead to the tilt of the actuators in the DM [Fig. 4(b)].

 figure: Fig. 4

Fig. 4 The temperature gradient maps and the displacement maps of the DM at ΔTe=5. The temperature gradient map (a) and the displacement map (b) without any distortion compensation. The temperature gradient map (c) and the displacement map (d) with the TCM working at the optimal compensation temperature.

Download Full Size | PDF

Accordingly, the whole surface shape (WSS) of the DM is distorted by the TID. Figures. 5(a1) and 5(b1) show the distorted WSSs (i.e. the TID) of the DM at ΔTe= −5 and ΔTe=5, respectively. The PV values of the two distorted WSSs are both 0.56 μm. According to the color temperature diagrams, the square distributed local distortion array could be identified clearly. The distribution of the local distortion is well matched with the actuators’ array, which could be identified by print-through of the actuators. The power spectral density (PSD) curves of the WSSs of the DM, illustrating the frequency characterization, are calculated [25,26] and shown in Figs. 5(a2) and 5(b2). The red lines represent the standard lines of the PSD curves [27], while the vertical dashed lines indicate the boundary between the low-frequency distortion and the high-frequency distortion [28]. As shown in Figs. 5(a2) and 5(b2), the PSD curves of the distorted WSSs locate above the standard line from low-frequency to high-frequency, which demonstrates the surface shape of the DM with TIDs are not qualified. The 3D maps of the distorted WSSs after high pass filtering (SSHF) the first 40 orders Zernike aberration of Figs. 5(a1) and 5(b1) are shown in Figs. 5(a3) and 5(b3) respectively, while the 2D maps are displayed in Figs. 5(a4) and 5(b4). The black circles marked in Fig. 5 represents the actuators’ position in the DM. As shown in Figs. 5(a4) and 5(b4), the distribution of the bulges consistent well with that of the actuators. A convexity and concavity appear on both sides of each actuator except for the central actuator in Figs. 5(a4) and 5(b4), which shows the typical characteristic of the dynamic HFD caused by the tilt of the actuators [8]. The dynamic HFD becomes more obvious with the distance from the actuator to the mirror center increasing, which could be explained by that the tilted angle θi,j [Eq. (8)] is positively related to the distance. The PV values of the aberration shown in Figs. 5(a4) and 5(b4) are both 0.18 μm. As the self-compensation residues illustrated in Figs. 5(a5) and 5(b5), the TID of the DM couldn’t be well corrected. Consequently, as the compensation result, the HFD will remain on the surface of the DM. The PV values of the fitting residues are both 0.22μm, which are almost equal to the PV values of the HFD shown in Figs. 5(a4) and 5(b4). By comparing the patterns in Figs. 5(a) and 5(b), the TID is opposite and symmetry when the temperature rises or falls of the same degrees.

 figure: Fig. 5

Fig. 5 The TIDs of the DM without temperature compensation. (a1) is the distorted WSS at  ΔTe=-5. (b1) is the distorted WSS at ΔTe=5. (a2) and (b2) are the PSD curves of the distorted WSSs. (a3) and (b3) are the 3D maps of the distorted SSHFs. (a4) and (b4) are the 2D maps of (a3) and (b3), respectively. (a5) and (b5) are the 2D maps of the fitting residues of the distorted WSSs in (a1) and (b1) respectively, while (a6) and (b6) are the 3D maps of the fitting residues.

Download Full Size | PDF

Figure 6(a) shows the changing laws of the PV values and the RMS values of the WSS and the SSHF when  ΔTe varies from −5 to 5 . The PV and RMS curves shown in Fig. 6(a) are origin-symmetric. Besides, the PV/RMS values of the WSS/SSHF increase linearly with the absolute value of  ΔTe. Specifically, the PV value of the WSS increases 0.12μm when the absolute value of  ΔTe increases 1, and the PV value of the SSHF increases 0.04μm. The relationship between the PV values of the TID and the expansion coefficient difference (Δα) of the mirror and the steel base is investigated at  ΔTe= 5 [Fig. 6(c)]. From the black line shown in Fig. 6(c), the PV value of the WSS increases linearly with the fitting slope 0.06 μm/(106/).

 figure: Fig. 6

Fig. 6 The surface shape distortion of the DM with (a) different temperature difference and (c)different thermal expansion coefficient difference at  ΔTe= 5. (b) and (d) are the self-compensation residues of (a) and (c).

Download Full Size | PDF

The TID of the DM depends on the thermal expansion coefficient difference and the temperature variation based on the curves in Fig. 6. As the thermal expansion coefficients of two different materials could hardly achieve the same value, the TID would always exist on the surface shape of a DM if there is temperature variation. As shown in Figs. 6(b) and 6(d), the PV values of the self-compensation residues are almost equal to the PV values of the SSHF, which are linearly increasing with the temperature and thermal expansion coefficient difference.

3.2 Depression of TID by using TCM on DM

To eliminate the TID, a TCM analyzed in section 2 is designed and installed on the steel base of the manufactured DM [Fig. 3]. In the simulation, the working temperature of the DM is also set ΔTe deviation from the original manufacturing environment temperature Tm. Meanwhile, the TCM is set to the temperature Ttcm and the temperature of the steel base is affected, while the temperature of the mirror still equals to the working environment temperature. Figures 4(c) and 4(d) show the temperature gradient map and the displacement map of the whole DM with the TCM working at the optimal compensation temperature. As shown in Fig. 4(c), when the temperature of the steal base is regulated by the TCM working at the optimal compensation temperature, ΔTeB is equal to 1.8, while ΔTeG is still equal to 5 (ΔTe). As the expansion mismatch is compensated, the tilt of the actuators is corrected [Fig. 4(d)]. Therefore, the TIDs shown in Figs. 5(a1) and 5(b1) turn into the patterns shown in Figs. 7(a1) and 7(b1) when ΔTe equals to −5 and 5, respectively.

 figure: Fig. 7

Fig. 7 The TIDs with the TCM working at the optimal compensation temperature. (a1) is the pre-compensated WSS at ΔTe= −5. (b1) is the pre-compensated WSS atΔTe=5. (a2) and (b2) are the PSD curves of the compensated WSSs in (a1) and (b1). (a3) and (b3) are the 3D maps of the pre-compensated SSHFs. (a4) and (b4) are the 2D maps of (a3) and (b3), respectively. (a5) and (b5) are the 2D maps of the fitting residues of the WSS in (a1) and (b1) respectively, while (a6) and (b6) are the 3D maps.

Download Full Size | PDF

The pre-compensated WSSs and the final fitting residuals are depicted in Figs. 7(a) and 7(b) respectively, when the temperature of the steel base is regulated to achieve the target ΔTeB [Eq. (12)] at ΔTe= −5 and ΔTe= 5. Note that in the pre-compensation process, only the TCM working to compensate the TID, while in the self-compensation process, the DM is controlled to realize the final correction. The PV values of the pre-compensated WSSs are both 0.10μm, as shown in Figs. 7(a1) and 7(b1). Different from Fig. 4, the convexity [Fig. 5(a2)] and concavity [Fig. 5(b2)] on both sides of each actuator no longer exist. All the convex hulls [Fig. 7(a1)] and the concave hulls [Fig. 7(b1)] on the surface shape move to the position of each actuator. Figures 7(a3) and 7(b3) show the 3D maps of the pre-compensated SSHFs after filtering the first 40 orders Zernike aberration of Figs. 7(a1) and 7(b1), while the 2D maps are shown in Figs. 7(a4) and 7(b4). The PV values of Figs. 7(a3) –7(a4) and 7(b3) –7(b4) are 0.08μm. Figures 7(a2) and 7(b2) display the PSD curves of the pre-compensated WSSs of the DM with the TCM working at the optimal compensation temperature. Although the PSD curve still locates above the standard line from low-frequency to high-frequency, the gap between the PSD curve and the standard line becomes smaller, comparing to Figs. 5(a2) and 5(b2).

Thus, by using the TCM, the distorted WSSs of the DM shown in Figs. 5(a1) and 5(b1) are improved to the pre-compensated WSSs shown in Figs. 7(a1) and 7(b1), which do not contain dynamic HFD caused by the actuators tilt. This indicates that incline angle θi,j reaches to zero [Eq. (10)] and the dynamic HFD caused by the shear components Δhijx and Δhijy [Eq. (6)] are eliminated. Therefore, only the normal component Δhijz [Eq. (6)], which could be corrected by the DM itself, remains on the surface shape. Based on the new surface shapes shown in Figs. 7(a1) and 7(b1), the DM could achieve better self-compensation results [Figs. 7(a5) and 7(b5)]. As shown in Figs. 7(a5) and 7(b5), the PV values of the fitting residuals of the DM self-compensation could achieve as good as 5nm, which indicates that the TID of the DM could be well depressed by using the TCM as pre-compensation method and the DM as the distortion-compensation approach.

Figures 8(a) and 8(b) show the relationship between the PV value of the WSS and the set compensation temperature TTCM of the TCM at different environment temperature. Correspondingly, the varying tendency of the PV value of the SSHF is illustrated in Figs. 8(c) and 8(d). As shown in Fig. 8, for different environment temperatures, there always exists an optimal compensation temperature TTCM, where the PV value of the WSS/SSHF could achieve the minimum. For instance, as the red curve shown in Fig. 8(a), the optimal compensation temperature of the TCM for the 15 environment temperature is 23.

 figure: Fig. 8

Fig. 8 The PV values of the TIDs with the TCM working at different temperatures. (a)–(d) show the influence of the set compensation temperature of the TCM at different Te. (e) shows the influence of the ΔTe on the optimal ΔTTCM, the ΔTeB and the PV of the WSS.

Download Full Size | PDF

As shown in Fig. 8(e), summarized from Figs. 8(a) and 8(b), the optimal ΔTTCM (blue solid circles) and ΔTeB (red solid circles) exist for each ΔTe. The optimal ΔTTCM is the difference between the optimal TCM temperature TTCM and the manufacturing temperature Tm. For instance, the optimal ΔTTCM and ΔTeB are 3 and −2 when ΔTe is equal to −5. The optimal ΔTeB is linear with the ΔTe, and the fitting slope is 0.36, which is almost the value of the αG/αB by introducing the material parameters listed in Table 1. The PV value (green hollow diamonds) of the WSS at the optimal ΔTeB is also linear with ΔTe (0.02μm/ fitting slope). As Fig. 8(e) shown, different from Figs. 6(b) and 6(d), the PV values of the self-compensation residue at the optimal ΔTeB are nearly equal to zero and stable. This indicate that the residual TID could be compensated by the DM itself even the remained normal component increases with ΔTe when the TCM working at the optimal temperature.

4. Experiments and results

The interferometric technique is the widely used to measure the optical surface shape. However, the accuracy of the interferometer is heavily influenced by the stability of the temperature. The deflectometry system (DS) [29,30] is chosen to measure the evolution process of the surface shape of the DM. As shown in Figs. 9(a) and 9(b), a DS consisting of a LCD display (1024 × 768 pixels, UM-900 of Lilliput Co., Ltd), a CCD camera (1384 × 1036 pixels, GS3-U3-14S5M-C of Point Grey Research Inc.) is built. The lab-manufactured 49-actuator DM (manufacture temperature Tm= 23) [Fig. 9(c)] with an 84mm × 84mm continuous mirror plate (>98%HR@1064nm and >50%R@632nm) and square distributed stacked PZT actuators (P885.91, Physik Instrumente GmbH, 5mm × 5mm × 36mm, 38μm stroke) array was rigidly connected to the steel base. The auxiliary TCM consisting of the TECs and a water-cooling block is screwed to the steel base as shown in Figs. 9(c) and 9(d). The material parameters of the steel base and the mirror are listed in Table 1.

 figure: Fig. 9

Fig. 9 (a) is the experiment setup; (b) is the layout of the experiment setup containing the control circuits; (c) is the lab-manufactured DM with a TCM; (d) is the bottom of the water-cooling block.

Download Full Size | PDF

Before measurement, the DS should be calibrated to acquire the relationship between the camera coordinate and the world coordinate [30]. After calibration, the reference phase of the measured DM is acquired, setting the ambient temperature equal to the mounting temperature (Tm). Two sets of grating patterns with sinusoidal intensity profile along the X and Y directions are displayed on the LCD and imaged on the CCD camera by the reflection of the surface of the DM [Fig. 9(a)]. To obtain high measurement accuracy, the LCD produces four phase-shift fringe images and the CCD captures four times repeatedly. A phase-shift algorithm is applied to computer the phase distribution on the DM. The measurement result is used as the reference phase, which reflects the original surface shape of the DM. When the ambient temperature is set to the target working temperature Te, the surface shape of the DM will change accordingly, which results in a new phase distribution. Then, the four phase-shift fringe images and algorithm are applied to computer the new phase distribution. Base on the reference phase, the phase difference of each pixel is calculated and the variation of the surface shape of the DM could be achieved using a reconstruct algorithm [27].

In the experiment, the TIDs of the DM without the TCM working are measured first. When the ambient temperature is set to 17.3 or 28.7 (ΔTe=-5.7, 5.7), the distorted WSSs of the DM are shown in Figs. 10(a1) and 10(b1). Figures 10(a2) and 10(b2) display the PSD curves of the distorted WSSs of the DM without the TCM working, at ΔTe=-5.7 and 5.7. As shown in Figs. 10(a2) and 10(b2), the PSD curves of the distorted WSSs locate above the standard line from low-frequency to high-frequency, which demonstrates the surface shape of the DM with TIDs are not qualified.

 figure: Fig. 10

Fig. 10 The TIDs of the DM without the TCM working. (a1) is the distorted WSS at ΔTe=-5.7, PV value = 2.17μm. (b1) is the distorted WSS at ΔTe=5.7, PV value = 1.97μm. (a2) and (b2) are the PSD curves of the WSSs. (a3) is the SSHF after filtering the first 40-orders Zernike aberration of (a1), PV value = 0.71μm, with the actuators’ position marked by black circles. (b3) is the SSHF after filtering the first 40-orders Zernike aberration of (b1), PV value = 0.69μm. with the actuators’ position marked by black circles. (a4) and (b4) are the influence of ΔTe on the PV and RMS of the distorted WSSs.

Download Full Size | PDF

The convex hulls [Fig. 10(a)] and the concave hulls [Fig. 10(b)] appear with an approximate square matrix distribution on the surface shape. Figures 10(a3) and 10(b3) show the SSHF of Figs. 10(a1) and 10(b1) with 7 × 7 square matrix black circles representing the actuators’ distribution, respectively. The convex and the concave hulls locate at both sides of the actuators [Figs. 10(a3) and 10(b3)]. Different from the simulation, in the experiment, due to the inhomogeneity of the assembly stress, not all the hulls could be observed on the surface shape of the DM. Figures 10(a4) and 10(b4) show that the PV and RMS values of the distorted WSSs and SSHFs increase linearly with the absolute value of ΔTe.

Figures 11(a1) –11(a2) and Figs. 11(b1) –11(b2) show the WSSs and the SSHFs of the compensation residue by the DM at ΔTe=-5.7 and 5.7 , respectively, without the TCM working. As shown in Fig. 11(a1), after compensation, the HFD still remain on the surface of the DM. Consequently, the convex and concave hulls appear more distinct in the compensation residue [Figs. 11(a1) and 11(b1)] than in the initial distorted WSSs [Figs. 10(a1) and 10(b1)]. At ΔTe= −5.7 , the PV values of the WSSs before and after compensation are 2.17μm [Fig. 10(a1)] and 1.46μm [Fig. 11(a1)], while the PV values of the SSHFs are 0.71μm [Fig. 10(a3)] and 1.02μm [Fig. 11(a2)]. At ΔTe= 5.7 , the PV values of the WSSs before and after compensation are 1.97μm [Fig. 10(b1)] and 1.41μm [Fig. 11(b1)], while the PV values of the SSHFs are 0.69μm [Fig. 10(b3)] and 0.75μm [Fig. 11(b2)]. Figures 11(a3) and 11(b3) show the PSD curves of the initial distorted WSSs [Figs. 10(a1) and 10(b1)] in black and the compensation residues without TCM working [Figs. 11(a1) and 11(b1)] in green. As shown in Figs. 11(a3) and 11(b3), the PSD curves of the compensation residues are almost the same with that of the initial distorted WSSs, which both locate above the standard line.

 figure: Fig. 11

Fig. 11 The compensation residues of the TIDs without the TCM working. (a1) is the compensation residue at ΔTe= −5.7 , PV = 1.46μm, RMS = 0.27 μm. (b1) is the compensation residue at ΔTe=5.7, PV = 1.41μm, RMS = 0.29μm. (a2) (PV = 1.02μm, RMS = 0.09μm) and (b2) (PV = 0.75μm, RMS = 0.08μm) are the SSHFs of the compensation residues after filtering the first 40-orders Zernike aberration. (a3) and (b3) are the PSD curves of the TIDs and the compensation residues without TCM working at ΔTe= −5.7 and ΔTe=5.7 , respectively.

Download Full Size | PDF

Experiment results above indicate that the TID (i.e. the distorted WSS) is the dynamic HFD, which could not be effectively compensated and will remain on the surface of the DM. Restricted by the phenomenon of the unwanted and uncorrectable TID, the correction capability of the DM is seriously limited in the practical working environment for the temperature variation.

A hybrid close-loop control algorithm (HCLA) is proposed to compensate the TID of the manufactured DM using an auxiliary TCM. As shown in Fig. 12, the HCLA consists of two major compensation steps, including the TCM pre-compensation and the self-compensation of the DM. In the TCM pre-compensation step, the steel base’s temperature is affected directly by the TCM, which is driven by the computer controlled TECs. Along with the temperature variation of the steel base, the TID is compensated accordingly and the HFD of the DM disappears from the surface shape of the DM. When the HFD corresponding to the actuators couldn’t be observed in the WSS of the DM, the self-compensation of the DM is triggered. In the self-compensation step, the normal displacement on the surface shape is compensated by the pushing and pulling activities of the actuators of the DM. Here, in our experiment, the PV value of the residue WSS (judgement value) is set as the ending criterion for the close-loop algorithm. If the judgement value reaches 0.5μm, the hybrid close-loop algorithm is accomplished and a well-compensated surface shape is finally achieved. If the judgement value is still higher than 0.5μm, the HFD caused by the shear displacements is considered remaining on the surface and the hybrid close-loop algorithm returns to the first step.

 figure: Fig. 12

Fig. 12 The hybrid close-loop control algorithm.

Download Full Size | PDF

In the experiment, ΔTeB of the steel base is controlled by the TCM and measured in real time. Figures 13(a1) and 13(b1) depict the pre-compensation results of the TIDs after the TCM compensation step while ΔTeB is measured −3.5 at ΔTe= −5.7 and 3.2 at ΔTe=5.7 . The WSSs of the TCM-compensation results are shown in Figs. 13(a1) and 13(b1) when ΔTe= −5.7 and 5.7 , respectively. The convex and the concave hulls in Fig. 10 almost disappear from the surface shape of the DM with the PV values of the WSSs decreasing from 2.17μm and 1.97μm [Figs. 10(a1) and 10(b1)] to 1.61μm and 1.48μm [Figs. 13(a1) and 13(b1)].

 figure: Fig. 13

Fig. 13 (a1) and (b1) are the WSSs of the pre-compensation results of the TIDs after the first step (only the TCM working) at ΔTe=-5.7 (PV = 1.61μm, RMS = 0.32μm) and ΔTe= 5.7 (PV = 1.48μm, RMS = 0.30μm). (a2) and (b2) are the self-compensation results when the hybrid close-loop algorithm is accomplished at ΔTe= −5.7 (PV = 0.44μm, RMS = 0.09μm) and ΔTe= 5.7 (PV = 0.45μm, RMS = 0.09μm). (a3) and (b3) are the SSHF of the self-compensation residues after filtering the first 40-orders Zernike aberration at ΔTe= −5.7 (PV = 0.22μm, RMS = 0.03μm) and ΔTe= 5.7 (PV = 0.22μm, RMS = 0.03μm). (a4) and (b4) are the PSDs of the initial distorted WSSs and the compensation residues after the hybrid close-loop algorithm is accomplished at ΔTe= −5.7 and ΔTe= 5.7.

Download Full Size | PDF

Figures 13(a2) –13(a3) and 13(b2) –13(b3) depict the self-compensation residues of the TID when the hybrid close-loop algorithm is accomplished. The WSSs of the residue at ΔTe= −5.7 and 5.7 are shown in Figs. 13(a2) and 13(b2). As seen from Figs. 13(a2) and 13(b2), no convex and concave hulls in square array distribution corresponding to the actuators could be observed. At ΔTe= −5.7 , the PV value of the WSS after self-compensation is 0.44μm [Fig. 13(a2)], only 20% of the initial distorted WSS [Fig. 10(a1)], while the PV value of the SSHF after self-compensation [Fig. 13(a3)] is 0.22μm, only 31% of the initial distorted SSHF [Fig. 10(a3)]. Correspondingly, at ΔTe= 5.7 , the PV value of the WSS after self-compensation [Fig. 13(b2)] is 0.45μm, which is 23% of the initial distorted WSS [Fig. 10(b1)], while the PV value of the SSHF after self-compensation [Fig. 13(b3)] is 0.21μm, which is 30% of the initial distorted SSHF [Fig. 10(b3)]. Figures 13(a4) and 13(b4) show the PSD curves of the initial distorted WSSs [Figs. 10(a1) and 10(b1)] in black and the self-compensation residues [Figs. 13(a2) and 13(b2)] in green when the hybrid close-loop algorithm is accomplished. As shown in Figs. 13(a4) and 13(b4), the PSD curves of the self-compensation residues shift away from the black curves and get very close to the red standard line, which demonstrates that the TIDs could be effectively compensated when the hybrid close-loop control is achieved.

Thus, by using the TCM, the initial TID of the DM shown in Figs. 10(a1) and 10(b1) are improved to the surface shapes shown in Figs. 13(a1) and 13(b1), which do not contain dynamic HFD caused by temperature variation. Based on the surface shapes shown in Figs. 13(a1) and 13(b1), a HCLA is carried out and a well-compensated WSS is finally achieved. As shown in Figs. 13(a2) and 13(b2), the PV values of the compensation residues of the DM could achieve as good as 23% of the initial TID. The application field of a DM is not restricted by the working temperature and could be extended to the working environment with a wider ambient temperature range.

In the experiment, some parameters are not as ideal as that set in the simulation, especially the influence function, the target temperature of the TCM and the residual assembly stress between the mirror and the 49-actuators. Thus, limited by the experimental condition, when the hybrid close-loop algorithm is accomplished in the experiment, the self-compensation residues are not as good as that in the simulation.

5. Conclusion

In conclusion, the mechanism and the compensation method of the TID of the DM is investigated theoretically and experimentally. The actuators’ tilt caused by the thermal expansion coefficient difference between the mirror and the steel base is the essential reason of the TID. Due to the tilt actuators’ shear components, the TID couldn’t be effectively corrected by the DM itself. To eliminate the actuators’ tilt, an auxiliary temperature module is proposed to regulate the steel base’s temperature. The compensation principle and the feasibility of the TCM are investigated. A finite element model is built to investigate the influence of the temperature and the thermal expansion coefficient differences on the TID and the TCM’s compensation capability. Simulation results show that the TID, positively related to the temperature and the thermal expansion coefficient difference, could be compensated effectively by the TCM. In experiment, a HCLA is adopted to control the compensation process of the TCM and the DM. Experiment results show that the TID is well depressed by using the TCM and the DM itself. Finally, a well-compensated DM surface shape is achieved in our experiment.

Funding

National Natural Science Foundation of China (NSFC) (61775112).

References

1. M. W. Ao, P. Yang, Z. P. Yang, E. D. Li, C. H. Rao, and W. H. Jiang, “A method of aberration measurement and correction for entire beam path of ICF beam path,” Proc. SPIE 6823, 68230I (2007). [CrossRef]  

2. B. Wattellier, J. Fuchs, J. P. Zou, K. Abdeli, H. Pépin, and C. Haefner, “Repetition rate increase and diffraction-limited focal spots for a nonthermal-equilibrium 100-TW Nd:glass laser chain by use of adaptive optics,” Opt. Lett. 29(21), 2494–2496 (2004). [CrossRef]   [PubMed]  

3. P. Yang, Y. Ning, X. Lei, B. Xu, X. Li, L. Dong, H. Yan, W. Liu, W. Jiang, L. Liu, C. Wang, X. Liang, and X. Tang, “Enhancement of the beam quality of non-uniform output slab laser amplifier with a 39-actuator rectangular piezoelectric deformable mirror,” Opt. Express 18(7), 7121–7130 (2010). [CrossRef]   [PubMed]  

4. S. Bonora, Y. Jian, P. Zhang, A. Zam, E. N. Pugh Jr., R. J. Zawadzki, and M. V. Sarunic, “Wavefront correction and high-resolution in vivo OCT imaging with an objective integrated multi-actuator adaptive lens,” Opt. Express 23(17), 21931–21941 (2015). [CrossRef]   [PubMed]  

5. S. Niu, J. Shen, C. Liang, Y. Zhang, and B. Li, “High-resolution retinal imaging with micro adaptive optics system,” Appl. Opt. 50(22), 4365–4375 (2011). [CrossRef]   [PubMed]  

6. J. Shamir, D. G. Crowe, and J. W. Beletic, “Improved compensation of atmospheric turbulence effects by multiple adaptive mirror systems,” Appl. Opt. 32(24), 4618–4628 (1993). [CrossRef]   [PubMed]  

7. M. L. Spaeth, K. R. Manes, C. C. Widmayer, W. H. Williams, P. K. Whitman, M. A. Henesian, I. F. Stowers, and J. Honig, “National Ignition Facility wavefront requirements and optical architecture,” Opt. Eng. 43(12), 2854–2865 (2004). [CrossRef]  

8. L. C. Sun, L. Huang, M. Yan, J. B. Fan, Y. M. Zheng, and C. Sun, “Simulational and experimental investigation on the dynamic high frequency aberration of the deformable mirror,” Opt. Express 25(26), 32853–32866 (2017). [CrossRef]  

9. L. Sun, Y. Zheng, C. Sun, and L. Huang, “Simulational and experimental investigation on the actuator-corresponding high-frequency aberration of the piezoelectric stacked array deformable mirror,” Opt. Express 26(18), 23613–23628 (2018). [CrossRef]   [PubMed]  

10. J. H. Lee, Y.-C. Lee, and E.-C. Kang, “A cooled deformable bimorph mirror for a high power laser,” J. Opt. Soc. Korea 10(2), 57–62 (2006). [CrossRef]  

11. C. Bruchmann, M. Appelfelder, E. Beckert, R. Eberhardt, and A. Tünnermann, “Thermo-mechanical properties of a deformable mirror with screenprinted actuator,” Proc. SPIE 8253, 82530D (2012). [CrossRef]  

12. K. A. Morse, S. L. McHugh, and J. Fixler, “Thermomechanical characterization of a membrane deformable mirror,” Appl. Opt. 47(29), 5325–5329 (2008). [CrossRef]   [PubMed]  

13. Z. Zhu, Y. Li, J. Chen, J. Ma, and J. Chu, “Development of a unimorph deformable mirror with water cooling,” Opt. Express 25(24), 29916–29926 (2017). [CrossRef]   [PubMed]  

14. Q. Xue, L. Huang, P. Yan, M. Gong, Z. Feng, Y. Qiu, T. Li, and G. Jin, “Research on the particular temperature-induced surface shape of a National Ignition Facility deformable mirror,” Appl. Opt. 52(2), 280–287 (2013). [CrossRef]   [PubMed]  

15. Q. Bian, L. Huang, X. K. Ma, Q. Xue, and M. L. Gong, “Effect of the particular temperature field on a National Ignition Facility deformable mirror,” Opt. Commun. 374, 119–126 (2016). [CrossRef]  

16. G. Vdovin, O. Soloviev, M. Loktev, S. Savenko, and L. Dziechciarczyk, “Optimal correction and feedforward control of low-order aberrations with piezoelectric and membrane deformable mirrors,” Proc. SPIE 8165, 81650W (2011). [CrossRef]  

17. R. Winsor, A. Sivaramakrishnan, and R. Makidon, “Low cost membrane type deformable mirror with high density actuator spacing,” Proc. SPIE 4007, 563–572 (2000). [CrossRef]  

18. J. Ma, Y. Liu, T. He, B. Li, and J. Chu, “Double drive modes unimorph deformable mirror for low-cost adaptive optics,” Appl. Opt. 50(29), 5647–5654 (2011). [CrossRef]   [PubMed]  

19. L. J. Hornbeck, “Deformable mirror spatial light modulators,” Proc. SPIE 1150, 1150III (1989).

20. L. Hu, L. Xuan, Y. Liu, Z. Cao, D. Li, and Q. Mu, “Phase-only liquid crystal spatial light modulator for wavefront correction with high precision,” Opt. Express 12(26), 6403–6409 (2004). [CrossRef]   [PubMed]  

21. B. P. Wallace, P. J. Hampton, C. H. Bradley, and R. Conan, “Evaluation of a MEMS deformable mirror for an adaptive optics test bench,” Opt. Express 14(22), 10132–10138 (2006). [CrossRef]   [PubMed]  

22. D. Guzmán, F. J. D. C. Juez, R. Myers, A. Guesalaga, and F. S. Lasheras, “Modeling a MEMS deformable mirror using non-parametric estimation techniques,” Opt. Express 18(20), 21356–21369 (2010). [CrossRef]   [PubMed]  

23. M. Tabatabaian, COMSOL for Engineers (Mercury Learning & Information, 2014).

24. R. W. Pryor, Multiphysics Modeling Using COMSOL: A First Principles Approach (Jones and Bartlett Publishers, 2011).

25. R. Z. Zhang, B. W. Cai, C. L. Yang, Q. Xu, and Y. Y. Gu, “Calculation of the power spectral density of the Wavefront,” Proc. SPIE 4231, 295–300 (2000). [CrossRef]  

26. J. M. Elson and J. M. Bennett, “Calculation of the power spectral density from surface profile data,” Appl. Opt. 34(1), 201–208 (1995). [CrossRef]   [PubMed]  

27. Draft International Standard ISO 10110 Part 8: Surface Texture (ISO@TC 172@SC 1@WG 2).

28. C. R. Wolfe and J. K. Lawson, “Measurement and analysis of wavefront structure from large-aperture ICF optics,” Office of Scientific & Technical Information Technical Reports 2633, 361–385 (1995).

29. P. Su, R. E. Parks, L. Wang, R. P. Angel, and J. H. Burge, “Software configurable optical test system: a computerized reverse Hartmann test,” Appl. Opt. 49(23), 4404–4412 (2010). [CrossRef]   [PubMed]  

30. L. Huang, C. L. Zhou, W. C. Zhao, H. Choi, L. Graves, and D. Kim, “Close-loop performance of a high precision deflectometry controlled deformable mirror (DCDM) unit for wavefront correction in adaptive optics system,” Opt. Commun. 393, 83–88 (2017). [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 (13)

Fig. 1
Fig. 1 (a) Simplified schematic of the NIF-type deformable mirror structure. (b) The XOY coordinates and the distribution of the actuators.
Fig. 2
Fig. 2 (a) The XYZ coordinates and the total displacements of the actuator. (b) The XOY and X ' O ' Y ' coordinates and the shear displacement of the mirror. (c) The Y ' O ' Z ' section and the normal displacement of the mirror.
Fig. 3
Fig. 3 (a)The DM with the TCM. (b) Distribution of four TECs.
Fig. 4
Fig. 4 The temperature gradient maps and the displacement maps of the DM at Δ T e =5. The temperature gradient map (a) and the displacement map (b) without any distortion compensation. The temperature gradient map (c) and the displacement map (d) with the TCM working at the optimal compensation temperature.
Fig. 5
Fig. 5 The TIDs of the DM without temperature compensation. (a1) is the distorted WSS at  Δ T e =-5. (b1) is the distorted WSS at  Δ T e =5. (a2) and (b2) are the PSD curves of the distorted WSSs. (a3) and (b3) are the 3D maps of the distorted SSHFs. (a4) and (b4) are the 2D maps of (a3) and (b3), respectively. (a5) and (b5) are the 2D maps of the fitting residues of the distorted WSSs in (a1) and (b1) respectively, while (a6) and (b6) are the 3D maps of the fitting residues.
Fig. 6
Fig. 6 The surface shape distortion of the DM with (a) different temperature difference and (c)different thermal expansion coefficient difference at  Δ T e = 5. (b) and (d) are the self-compensation residues of (a) and (c).
Fig. 7
Fig. 7 The TIDs with the TCM working at the optimal compensation temperature. (a1) is the pre-compensated WSS at Δ T e = −5. (b1) is the pre-compensated WSS at Δ T e =5. (a2) and (b2) are the PSD curves of the compensated WSSs in (a1) and (b1). (a3) and (b3) are the 3D maps of the pre-compensated SSHFs. (a4) and (b4) are the 2D maps of (a3) and (b3), respectively. (a5) and (b5) are the 2D maps of the fitting residues of the WSS in (a1) and (b1) respectively, while (a6) and (b6) are the 3D maps.
Fig. 8
Fig. 8 The PV values of the TIDs with the TCM working at different temperatures. (a)–(d) show the influence of the set compensation temperature of the TCM at different T e . (e) shows the influence of the Δ T e on the optimal Δ T TCM , the Δ T eB and the PV of the WSS.
Fig. 9
Fig. 9 (a) is the experiment setup; (b) is the layout of the experiment setup containing the control circuits; (c) is the lab-manufactured DM with a TCM; (d) is the bottom of the water-cooling block.
Fig. 10
Fig. 10 The TIDs of the DM without the TCM working. (a1) is the distorted WSS at Δ T e =-5.7, PV value = 2.17μm. (b1) is the distorted WSS at Δ T e =5.7, PV value = 1.97μm. (a2) and (b2) are the PSD curves of the WSSs. (a3) is the SSHF after filtering the first 40-orders Zernike aberration of (a1), PV value = 0.71μm, with the actuators’ position marked by black circles. (b3) is the SSHF after filtering the first 40-orders Zernike aberration of (b1), PV value = 0.69μm. with the actuators’ position marked by black circles. (a4) and (b4) are the influence of Δ T e on the PV and RMS of the distorted WSSs.
Fig. 11
Fig. 11 The compensation residues of the TIDs without the TCM working. (a1) is the compensation residue at Δ T e = −5.7 , PV = 1.46μm, RMS = 0.27 μm. (b1) is the compensation residue at Δ T e =5.7, PV = 1.41μm, RMS = 0.29 μm. (a2) (PV = 1.02μm, RMS = 0.09 μm) and (b2) (PV = 0.75μm, RMS = 0.08 μm) are the SSHFs of the compensation residues after filtering the first 40-orders Zernike aberration. (a3) and (b3) are the PSD curves of the TIDs and the compensation residues without TCM working at Δ T e = −5.7 and Δ T e =5.7 , respectively.
Fig. 12
Fig. 12 The hybrid close-loop control algorithm.
Fig. 13
Fig. 13 (a1) and (b1) are the WSSs of the pre-compensation results of the TIDs after the first step (only the TCM working) at Δ T e =-5.7 (PV = 1.61 μm, RMS = 0.32 μm) and Δ T e = 5.7 (PV = 1.48 μm, RMS = 0.30 μm). (a2) and (b2) are the self-compensation results when the hybrid close-loop algorithm is accomplished at Δ T e = −5.7 (PV = 0.44 μm, RMS = 0.09 μm) and Δ T e = 5.7 (PV = 0.45 μm, RMS = 0.09 μm). (a3) and (b3) are the SSHF of the self-compensation residues after filtering the first 40-orders Zernike aberration at Δ T e = −5.7 (PV = 0.22μm, RMS = 0.03 μm) and Δ T e = 5.7 (PV = 0.22μm, RMS = 0.03 μm). (a4) and (b4) are the PSDs of the initial distorted WSSs and the compensation residues after the hybrid close-loop algorithm is accomplished at Δ T e = −5.7 and Δ T e = 5.7.

Tables (1)

Tables Icon

Table 1 Material parameters in the finite element simulation

Equations (13)

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

δ iA = x iA ' x i =Δ T e ia α G
δ jA = y jA ' y j =Δ T e ja α G
δ iB = x iB ' x i =Δ T e ia α B
δ jB = y jB ' y j =Δ T e ja α B
Δ T e = T e T m
{ Δ h ijz =Δ h ij cos( θ i,j ) Δ h ijx =Δ h ij sin( θ i,j )sin( α i,j ) Δ h ijy =Δ h ij sin( θ i,j )cos( α i,j )
tan( α i,j )= δ iA δ jA = i j (j0)
tan( θ i,j )= δ Bi,j δ Ai,j h
δ Ai,j = δ iA 2 + δ jA 2
tan( θ i,j )= a i 2 + j 2 (Δ T eB α B Δ T eG α G ) h
Δ T eB = α G α B Δ T eG
Δ T eB = α G α B Δ T e
Δ T eB =f( T e , T TCM , T m )
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.