Vous êtes sur la page 1sur 21

header for SPIE use

Comparison of stress relief procedures for cryogenic aluminum mirrors


Raymond G. Ohl*a, Michael P. Barthelmya, S. Wahid Zewaria, Ronald W. Tolanda, Joseph C. McMannb, David F. Pucketta, John G. Hagopiana, Jason E. Hylanc, J. Eric Mentzella, Ronald G. Minka, Leroy M. Sparra, Matthew A. Greenhousea, and John W. MacKentyd
b

NASA/Goddard Space Flight Center, Greenbelt, Md. ManTech International Corp., NASA/Goddard Space Flight Center, Greenbelt, Md. c Swales Aerospace, Inc., 5050 Powder Mill Road, Beltsville, Md. d Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, Md.
ABSTRACT

The Infrared Multi-Object Spectrograph (IRMOS) is a facility instrument for the Kitt Peak National Observatory Mayall Telescope (3.8 meter). IRMOS is a low- to mid-resolution (R = / = 3003000), near-IR (0.82.5 m) spectrograph that produces simultaneous spectra of ~100 objects in its 2.8 2.0 arcmin field of view using a commercial MEMS multi-mirror array device. The instrument operating temperature is ~80 K and the design is athermal. The optical bench and mirrors are machined from aluminum 6061-T651. In spite of its baseline mechanical stress relief, aluminum 6061-T651 harbors some residual stress, which, unless relieved during fabrication, may distort mirror figure to unacceptable levels at the operating temperature (~80 K). Other cryogenic instruments using aluminum mirrors for both ground-based and space IR astronomy have employed a variety of heat treatment formulae, with mixed results. We present the results of a test program designed to empirically determine the best stress relief procedure for the IRMOS mirrors. Identical test mirrors with spherical and flat optical prescriptions are processed with five different stress relief formulae from the literature and institutional heritage and compared to samples without any additional processing. After figuring via diamond turning, the mirrors are tested for figure error and radius of curvature at room temperature and at ~80 K for three thermal cycles. The heat treatment procedure for the mirrors that yielded the least and most repeatable change in figure error is applied to the IRMOS mirror blanks. We correlate the results of our optical testing with heat treatment and metallographic data. Keywords: mirrors, cryogenic, aluminum, heat treatment, stress relief, optomechanical design, IRMOS

1. INTRODUCTION
The Infrared Multi-Object Spectrograph (IRMOS) is a facility instrument for the Kitt Peak National Observatory (KPNO) Mayall Telescope (3.8 meter) that will see first light in the spring of 2003.1 The project is a collaboration of NASA/Goddard Space Flight Center (GSFC), the Space Telescope Science Institute (STScI), and KPNO. IRMOS is a low- to mid-resolution (R = / = 3003000), near-IR (0.82.5 m) spectrograph which produces simultaneous spectra of ~100 objects in its 2.8 2.0 arcmin field of view. The instrument uses a Texas Instruments, Inc. micro-electromechanical system (MEMS) multimirror array (MMA) device as a real-time programmable slit mask. The spectrograph operating temperature is ~80 K and the design is athermal: The optical bench and mirrors are machined from aluminum (Al) 6061-T651. The instrument optical design is an asymmetric arrangement of powered, off-axis aspheres and flat fold mirrors. The mirrors have an aspect ratio of ~6:1. Most mirrors have apertures ~100 100 mm in size, except for one that is 286 269 mm. The mirrors are mounted to the optical bench via a semi-kinematic, integral flexure mounting scheme.2,3 The fabrication specifications on mirror prescription are 1% for radius and 0.30.8% for conic constant. Surface error fabrication tolerances are as follows: figure error of < 0.1 RMS ( = 632.8 nm) at 293 K, where figure is roughly defined as surface error with spatial periods D 1 mm (i.e., deviation from the best-fit prescription, detectable with an interferometric null
*

Correspondence: Email: Raymond.G.Ohl@gsfc.nasa.gov; Tel.: 301-286-8368; Fax: 301-286-0204 Texas Instruments, Inc., 12500 TI Boulevard, Dallas, TX 75243-4136; Tel.: 800-336-5236

(a)

(b)

Figure 1. (a) Mirror aperture design. Note racetrack aperture corners. (b) Schematic showing where the test mirror blanks (squares) and dog bone tensile samples (thin rectangles) are removed from 1.2 m 1.2 m 25 mm Al plate. All samples are removed from the center of the plates thickness (the direction out of the page) and > 200 mm away from the edges in order to avoid areas with the potential for a greater and/or non-uniform stress pattern. The mirror blanks for each process are marked (spheres as n, flats as nA). The rolling/grain direction for the plate runs top-bottom in the plane of the page.

test) and microroughness of < 10 nm RMS, where microroughness is surface error with 1 D 100 m (i.e., detectable with a white light interferometer). These tolerances represent the best achievable surface error for state of the art single point diamond machining of Al 6061-T651 aspheres.3 In spite of its baseline mechanical stress relief, Al 6061-T651 harbors some residual stress and machining can introduce stress to the mirror blanks. The premise of this work is that residual stress in the mirror substrate combines with thermal stresses (i.e., temperature change) to change the figure error in an unanticipated manner at the operating temperature. The solution is to relieve stress during fabrication such that the change in figure from room temperature (293 K) to the operating temperature (80 K) is minimized and any measured change is repeatable with multiple thermal cycles. Other workers give a broad introduction to dimensional stability for metal mirrors47 and discuss specific stress relief treatments for Al alloys.813 We investigate stress relief (SR) processes in order to find one that achieves stability superior to straight Al 6061-T651 with little or no loss of mechanical properties. Mirror substrate strength is important because the mirror mount flexures must function per design. Other cryogenic instruments using Al mirrors for both ground-based and space IR astronomy have employed a variety of SR formulae, with varied results.1423 The literature is unclear on the best SR recipe. Experience from instrument programs at GSFC and KPNO is mixed. We present the results of a test program designed to empirically determine the best SR procedure for our instrument mirrors. Sets of identical, IRMOS-like test mirrors are processed with five different heat treatment formulae from the literature and GSFC heritage and compared to samples with out any additional processing (straight Al 6061-T651), for a total of six sets of mirrors. Each set consists of one mirror with flat and one with spherical optical prescription. After figuring via diamond turning, the mirrors are tested for figure error and radius of curvature change for three thermal cycles (293 K to 80 K). The SR procedure for the mirrors that yielded the least and most repeatable change in figure error is applied to the IRMOS mirror blanks. We correlate the results of the optical testing with heat treatment and metallographic data and draw conclusions regarding the physics of dimensional stability and SR procedures for Al 6061.

(a)

(b)

Figure 2. Surface plots of the figure error map for flat (a) and spherical (b) test mirrors for SR 1 after diamond turning (293 K). The magnitude of the surface error is 0.103 and 0.237 RMS ( = 632.8 nm), respectively. The 100 mm diameter, collimated test beam did not completely fill and was not well centered on the aperture of the flat (a). For the sphere (b), the full aperture is shown. The general character of the figure is typical of all of the test mirrors after fabrication (prior to cryogenic testing), but its amplitude is generally lower for the other mirrors. The peak/divot in the center of the array is probably caused by misalignment of the diamond tool with respect to the axis of rotation of the spindle.

2. MIRROR DESIGN AND FABRICATION


2.1. Mirror optomechanical design and fabrication Two test mirrors were fabricated for each SR process, one with flat and one with spherical prescription, because both flat and powered mirrors with similar aperture, aspect ratio, and radius are required by the IRMOS optomechanical design.1,3 The test mirrors have a 94 100 mm aperture (Figure 1a).24 The flat mirrors are 17.3 mm thick. The spherical mirrors have a radius of 400 mm (concave) and are cut such that they would be 22.9 mm thick at the corners of the aperture, were the corners sharp and not racetrack. The mirrors do not have mounting features --- they are simple, monolithic substrates.

The flat mirrors front and rear surfaces are highly parallel to the surfaces of the original Al plate stock (see below), while the sag of the spherical mirror increases the likelihood of inducing an asymmetric stress pattern during machining the nonuniform thickness. The blanks were rough-cut from a piece of commercial Al 6061-T651, 1.2 m 1.2 m 25 mm plate stock (Figure 1b).* Each blank was marked with a serial number indicating its original position and orientation in the plate. The blanks were removed from the center of the plate stock (in depth) to balance the best available grain structure. Also, the blanks were located away from the sides of the plate to minimize potential edge effects (i.e., to avoid a potential non-homogeneous stress pattern). Furthermore, the overall location of each blank in the plate was similar vis--vis distance from the sides and grain direction. Rough and finish machining of the mirror blanks was performed at Toper Manufacturing Company, Inc. The SR procedure was executed during fabrication by the Materials Engineering Branch laboratory at GSFC (Code 541). The optical surface fabrication specifications were a figure error of < 0.1 RMS ( = 632.8 nm), a microroughness of < 10 nm RMS, and a radius tolerance of 1% for the spheres. The mirrors were figured via single point diamond turning (SPDT) by Janos Technology, Inc. For the spheres and for flat SR 5, the vendor lapped the rear surface flat to ~0.5 P-V for direct mounting to the SPDT machines vacuum chuck. However, for flats SR 14 and 6, the vendor attached the blank to the spindle using epoxy and an intermediate plate fixture (this was thought to avoid adding stress to the blanks via the lapping

Alcoa lot number 718-982; Alcoa Corp., Alcoa Corporate Center, 201 Isabella Street, Pittsburgh, PA 15212-5858, USA; Tel.: 412-553-4545 Toper Manufacturing Company, Inc., 6427 Baltimore National Pike, Baltimore, MD 21228-3904, USA; Tel.: 410-7449110 Janos Technology, Inc., 1068 Grafton Road, Townshend, Vermont 05353-7702, USA; Tel.: 802-365-7714

process, but this was probably an unnecessary step). Surface plots of the figure error map for a typical flat and sphere after fabrication and prior to cryogenic testing are shown in Figure 2.
2.2. Material selection We considered three different types of Al 6061-T651 stock for these mirrors: extruded, forging, and plate stock. For extruded material, one would remove the mirror blanks such that the normal to the optical surface would be approximately parallel to the axis of the extrusion. Since the grains would be oriented perpendicular to the optical surface, this might yield mirrors with a low-amplitude, homogenous microroughness. However, one of the IRMOS mirrors is too large for the largest commercially available extrusion. Also, the presence of porosity becomes more likely at the center of thick or large diameter products.

We also considered removing the blanks from disk-like, pancake forgings, with the optical surface parallel to the plane of the disk. This would have produced mirrors with very homogeneous small-scale roughness (Sections 3.5 and 4.3). However, forgings involve significant lead-time and higher cost per blank compared to plate stock. For IRMOS, the impact of the character of the microroughness and the resulting scatter is an effect that is important to measure, but its control is a secondary consideration.25 Nevertheless, we fabricated a set of test mirrors from forgings via SR 4.* We compare their grain structure to the mirrors from plate in Section 4.3. We present their cryogenic performance elsewhere.26 We choose to use plate stock for the IRMOS mirrors and this test program, because it is readily available and inexpensive and encompasses the largest IRMOS mirror. The residual stress patterns associated with the formation of heat treatable Al alloy plate, extrusions, and forgings vary with product form and product thickness or diameter.27,28 For the purposes of interpreting the results of this experiment, we assume that the relative performance of each SR procedure for plate will be similar for other types of stock.
2.3. Candidate stress relief processes We investigate five candidate SR processes (Table 1). The procedures are based on metallurgical theory and literature, GSFC heritage, and the recommendations of an external optical materials consultant.29,30 These procedures are representative of the range of formulae discussed in the literature.823 Detailed procedures for the execution of these SR processes are given elsewhere.31 SR 1: These mirror blanks are cut from the Al 6061-T651 plate and received no additional thermal treatment. They represent a control. SR 2: This process is an overage stress relief heat treatment performed for 2 hours at 260 C. SR 2 will produce a lower residual stress level than the 177200 C stress relieving temperatures that are more commonly used on 6061-T6, but will also significantly degrade yield and ultimate strengths. SR 3: This is SR 2 with the addition of an up-hill quench.32 The latter treatment has been used by many workers for cryogenic Al mirrors.14,20,21 SR 4: This treatment uses a quench in a 28% UCON A glycol solution at 90 C.29 The glycol quench is established processing per AMS 2770.33 This SR has been widely used by those desiring to improve dimensional stability.34 SR 5: This is SR 4 with an uphill quench from liquid nitrogen (LN2) into boiling water, performed < 30 min following the solution treatment quench. This should represent the optimal conditions for uphill quenching because the blanks residual stress pattern is symmetric and its microyield strength is at a minimum. SR 6: This process is identical to SR 5 except that the 28% glycol 90 C quench is replaced by an agitated room temperature water quench. The latter is established processing for heat treatable Al alloys, but will generate higher residual stresses.

The following overview describes Al processing in general and our processing steps and terminology in greater detail:
Rough machining is defined as the oversized extraction of the proto-mirror blank from the plate (centered on the centerline of the thickness of the plate to 1 mm), trimming that piece to the final aperture size, and, finally, symmetrically removing material about the centerline from the rear and front of the blank until its thickness is oversized by 1.27 mm on the front and rear sides. Finish machining is defined as removing material from the front and rear of the blank in a symmetric fashion about the center line to meet final dimensions, but leaving 0.15 +/-0.10/0.00 mm material on the optical surface for SPDT.
*

Falcon Stainless and Alloys Corp., 39 Hawson Avenue, Waldwick, NJ 07463, USA; Tel.: 201-670-8300

Strengthening of Al alloys by heat treatment is accomplished by a precipitation hardening mechanism that involves two basic steps: solution treating and age hardening. In solution treating, the part is heated to a temperature approaching the melting point (530 C versus 582/652 C, respectively, for 6061). Since the solubility of the Mg and Si alloying additions in 6061 increases with temperature, they will completely or nearly completely dissolve in the Al matrix during the solution treatment soak period. Quenching the part at a sufficiently high rate results in a supersaturated solution of Mg and Si in Al. The second step, age hardening, may be performed at room or elevated temperature and allows diffusion of the alloying elements to form a coherent submicroscopic precipitate that obstructs lattice motion and, as a result, increases strength. The precipitation hardening process produces an excellent combination of mechanical properties in heat treatable Al alloys, but also introduces undesirable residual stresses --- primarily due to quenching. The main residual stress reduction approach used by producers is to mechanically stress relieve the product immediately following solution treatment, before natural aging can result in strengthening. Long products are stretched plastically several percent while shorter products are compressed. These are designated respectively as the Tx51 and Tx52 tempers for plate, extruded, and rolled shapes and forgings. Mechanically stress relieved sheet is identified by minimum residual stress or chem milling grade designations. Mechanical stress relief tends to be a very effective process, and numerous mirrors have been manufactured from this type of stock with additional low temperature stress relief and thermal cycling treatments applied. A second residual stress reduction approach is to tailor the quench process to produce a lower cooling rate. This can span sophisticated quench factor analysis approaches where cooling rate is optimized in several media to simpler steps such as substituting hot water or glycol/water mixes for the room temperature water bath. Mechanical properties and corrosion resistance limit the cooling rate reduction. Quenching parts into room temperature water can have devastating effects on machining dimensional stability, as evidenced by several GSFC experiences with Al 7075. Per Table 1, our glycol quench consisted of a 28% (volume) solution of Tenaxol Inc.s UCON A in water, held at 90 C,* with mild agitation during quenching. While 28% exceeds the 22% maximum listed in AMS 2770 for 25 mm sections of 6061 material, the effect of slightly lower cooling rates on mechanical properties in 6061 is small. Residual stress reduction steps that are supplementary to the two main approaches include uphill quenching, stress relieving, and thermal cycling. Uphill quenching is intended to neutralize the residual stress pattern of the downhill quench from the solution treatment temperature by quenching from a LN2 bath to a steam blast or boiling water bath. Our uphill quenching was performed by cooling thoroughly in a LN2 bath followed by quenching into a boiling water bath of sufficient size to prevent the temperature from dropping more than 6 C. Lack of facilities at GSFC precluded the use of steam.
Stress relieving Al alloys typically involves temperatures of 177200 C, which offer only modest stress relief and often slightly increase the strength of 6061. Higher stress relieving temperatures occasionally are used and reduce residual stresses to lower levels, but result in the loss of strength. The Cassini/Composite Infrared Spectrometer (CIRS) project at GSFC used 260 C in an attempt to minimize figure error change observed during thermal cycling.1719 After this treatment and optical testing over many thermal cycles, the CIRS mirrors met their figure error specification at their operating temperature. The change in CIRS mirror figure from room to operating temperature is similar to the SR 2 mirror change in this study (Section 3.4). This procedure also was used for the Space Infrared Telescope Facility/Infrared Array Camera (IRAC) mirrors.22 Thermal cycling stabilizes components by repeatedly alternating the part temperature between approximately 190 C and +150 C at a controlled rate of a few degrees C per minute. This obtains a measure of stress relaxation for those small stresses introduced in the final machining and finishing processes. A thermal cycling temperature range of 190 C to +150 C is employed in this study following finish machining/stress relief and following diamond turning and polishing, to precondition each part and lower residual stresses. This processing is easily performed in programmable thermal cycling chambers. We constructed a sealed, gaseous N2 purge chamber to protect the IRMOS mirrors from water condensation during thermal cycling.

In order to minimize thermal gradients that might distort parts, age hardening, stress relieving, and thermal cycling heat treatments all begin with the mirrors at room temperature. Heating and cooling rates are limited to 3 C per minute for aging and stress relieving and 2 C per minute for thermal cycling.

Tenaxol, Inc., 5801 W. National Ave., Milwaukee, WI 53214, USA; Tel.: 414-476-1400

Step

SR 1: Al 6061T651 (control)

SR 2: Simple, overage heat treatment

SR 3: SR 2 with uphill quench

SR 4: Solution treat using 28% glycol quench

SR 5: SR 4 with up-hill quench

SR 6: Solution treat with H2O quench and up-hill quench

1 2 3

Rough machine

Rough machine Heat treat at 260 C for 2 hours

Rough machine Heat treat at 260 C for 2 hours

Rough machine

Rough machine

Rough machine

Solution treatment at 530 C Quench within 15 sec in 28% UCON Quenchant A at 2935 C Uphill quench: Allow to reach 23 C, slowly place in LN2, dunk in boiling H2O Age at 175 C, 8 hrs Finish machine Finish machine Finish machine Finish machine Age at 175 C, 8 hrs TC 3 times at rates not exceeding 1.7 C/min, cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C SPDT and polish SPDT and polish TC 3 times at rates not exceeding 1.7 C/min; cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C Figure test at 293 K and 80 K, 3 TCs Figure test at 293 K and 80 K, 3 TCs Figure test at 293 K and 80 K, 3 TCs SPDT and polish SPDT and polish TC 3 times at rates not exceeding 1.7 C/min; cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C Figure test at 293 K and 80 K, 3 TCs

Solution treatment at 530 C Quench within 15 sec in 28% UCON Quenchant A at 2935 C Uphill quench: Allow to reach 23 C, slowly place in LN2, dunk in boiling H2O Age at 175 C, 8 hrs Finish machine Age at 175 C, 8 hrs TC 3 times at rates not exceeding 1.7 C/min, cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C SPDT and polish TC 3 times at rates not exceeding 1.7 C/min; cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C Figure test at 293 K and 80 K, 3 TCs

Solution treatment at 530 C Quench within 15 sec into H2O at 1824 C Uphill quench: Allow to reach 23 C, slowly place in LN2, dunk in boiling H2O Age at 175 C, 8 hrs Finish machine Age at 175 C, 8 hrs TC 3 times at rates not exceeding 1.7 C/min, cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C SPDT and polish TC 3 times at rates not exceeding 1.7 C/min; cool to 83 K, hold for 30 min, heat to 23 C, hold for 15 min, heat to 150 C, hold 30 min, cool to 23 C Figure test at 293 K and 80 K, 3 TCs

6 7 8

10

11

12

Table 1. Summary of candidate stress relief procedures (Section 2.3). The acronyms SR, TC, and SPDT stand for stress relief, thermal cycle, and single point diamond turning, respectively.

3. OPTICAL TESTING
Optical testing is performed in the Optics Branch laboratories at GSFC (Code 551).
3.1. Figure and radius testing 3.1.1. Experimental setup and procedure for flat mirrors Figure error is measured using a horizontal, single-pass, interferometric null test setup that incorporates a dewar with a large optical window that sits on a rail (Figure 3). The phase-shifting, Fizeau interferometer is a Zygo Corp. GPI-6XP512 with a 100 mm diameter reference flat that did not cover the full mirror aperture for the flat mirrors (i.e., the tips of the corners where excluded).* The dewar has a 300 mm diameter, 310 mm long, cylindrical, Al vacuum shroud.

The flange at the front of the dewar incorporates a through hole and a ~150 mm diameter, ~25 mm thick, plano-plano, fused silica optical window. The flange at the rear of the dewar supports an oxygen free high conductivity Cu rod (cold finger) that protrudes through the flange and 190 mm into the dewar. An Al plate bolted to the end of the cold finger hosts two long steel pins on which the mirror under test is simply supported. Figure tests of the mirrors on and off this mount indicate that these low-aspect ratio mirrors are not measurably distorted due to gravity sag. For better heat conduction between the mirror and cold plate, a 5 5 cm, <1.25 mm thick sheet of indium is sandwiched between the mirror and the cold plate, roughly centered on the mirror aperture. The mirror is also taped to the cold plate with Al or Cu tape in order to help keep it from separating from the cold finger during handling and improve conduction (Figure 4a). The temperature of the Al radiation shroud lags that of the cold finger by many hours (i.e., it does not get below ~250 K during a typical ~6 hour cool-down). In order to shield the sides of the mirror from radiation from this surface, we bolt a small, Cu sheet metal shroud with exterior thermal blanketing to the cold plate (Figure 4b). This reduces the mirrors exposure to warm solid angle and decreases the potential for a radial thermal gradient in the mirror. Modeling shows that this step was probably unnecessary (Section 3.2). We built a custom iris shutter and mounted it between the mirror and the dewar window (Figure 4c; see Section 3.2 for a discussion of its purpose). The shutter housing is Al and the blades of the iris aperture are steel. It is actuated using a Phytron, Inc. cryogenic stepper motor and controller. The dewar is bolted to a tip/tilt plate for alignment to the interferometer. The tip/tilt plate is bolted to a carrier that rides on a rail. The dewar accommodates four cryogenic, silicon diodes (Lake Shore Cryotronics, Inc. model DT-471). Temperature readout is via Lake Shore Cryotronics, Inc. model numbers 330 and 280. Temperature data were recorded by hand. We used four-wire diodes that have an accuracy of 0.1 C, but better precision. The diodes are attached to the cold finger, shutter housing, and side of the mirror in two places in order to measure a possible temperature gradient through its thickness (one near the front optical surface, one near the rear of the substrate; Figure 4a). The wire leads to the diodes are taped to the side of the cold finger to eliminate parasitic heat input to the mirror and ensure reliable measurements. The cryogenic figure test procedure is as follows: 1. Mount the mirror in the dewar as described above with the inscribed part number on the side of the substrate facing the ceiling (consistent with the orientation of the other test mirrors). Mount the dewar such that the mirror aperture is centered on the 100 mm diameter beam. 2. Measure the mirrors figure error through the dewar window. 3. Evacuate the dewar to 810-4 torr. 4. Measure the figure error.

Zygo Corporation, Laurel Brook Road, Middlefield, CT 06455-0448, USA; Tel.: 860-347-8506 Janis Research Company, Inc., 2 Jewel Drive, P.O. Box 696, Wilmington, MA 01887-0696, USA; Tel.: 978-657-8750 Phytron, Inc. 1347 Main Street, Waltham, MA 02451, USA; Tel.: 781-647-3581 Lake Shore Cryotronics, Inc., 575 McCorkle Blvd., Westerville, OH 43082, USA; Tel.: 614-891-2244

Flow LN2 through the exposed end of the cold finger such that it cools at the rate of ~100 K / hour and the mirror reaches a final temperature of 80 K. Record the temperature data from the diodes at regular intervals. The temperature gradient through the thickness of the mirror is never greater than about 0.4 C. 6. When the mirror reaches ~80 K, let the mirror approach equilibrium for at least one hour (i.e., when the rate of mirror temperature change is less than 0.1 K / 15 min for one hour, it is at equilibrium). 7. Measure the figure error at equilibrium. 8. Close the shutter and the mirror will further cool to a slightly lower equilibrium temperature. 9. At the new equilibrium temperature, open the shutter and quickly (i.e., within 30 sec) measure the figure error. 10. Halt the flow of LN2. Allow the dewar warm to 293 K. 11. When the mirror has returned to ~293 K, one thermal cycle is complete. Repeat the procedure, starting at step 4 above, for a total of three thermal cycles per mirror. The figure error data from steps 4 (293 K) and 9 (80 K) are compared to obtain the change in figure from 293 to 80 K (Section 3.3). This procedure was the norm for flat mirrors, with significant changes for spherical mirrors (Section 3.1.2) and/or when dewar window surface error was also measured (Section 3.2). This setup and procedure afforded a measurement precision of better than ~0.005 in RMS figure error in sampling the change in figure error from 293 to 80 K.
3.1.2. Experimental setup and procedure for spherical mirrors The figure test setup and procedure for the spherical mirrors is similar to that described for the flat mirrors (Section 3.1.1). We use a f/3.3 transmission sphere for the null test. The test beam covered the full aperture of the spherical mirror.

5.

In addition to figure error measurements, we used the rail and the interferometer to measure the change in radius of curvature over the course of the thermal cycle.35 To measure the radius of curvature at one temperature, we record the position of the carrier on the rail at the time of the figure error measurement. We then translate the mirror in focus toward the interferometer and measure the wavefront from the cats eye or retro-reflecting return from the surface of the mirror and record the associated rail position. The difference in these two rail positions (mirror figure measurement and cats eye) gives the radius of the mirror (at one temperature). In order to compare radius measurements at different temperatures, the mirror must remain stationary on the carrier. However, the cold finger and other dewar hardware connect the mirror to the carrier. When the cold finger cools, it contracts and moves the mirror roughly parallel to the rail. To remove this effect, we also measure the cats eye return from a surface of the dewar window. If f is the position of the carrier on the rail during a figure error measurement, m is its position in the cats eye configuration, and w is its position when the dewar window surface is in the cats eye configuration, then the change in radius from 293 to 80 K is given by

R 293 80 = f 293 f 80 ( w293 w80) (m 293 m80)


where the subscripts 293 and 80 denote warm and cold temperatures, respectively.

(1)

There are several sources of statistical and systematic error for this measurement. The alignment of the rail relative to the axis of the interferometer is off by < 2 mm, which produces a cosine error of ~0.001%. Deviation from the ideal performance of the rail (backlash, etc.) contributes to measurement error (~10 m). To determine the focus of the cats eye return, we bound the caustic visually by watching for defocus in the interferogram. This could be done in a more precise manner by using software to fit the power in the measured wavefront error. For mirror figure error measurements, the focus term from a least squares fit is minimized to < 0.1 peak-to-valley (P-V). The dewar hardware connecting the mirror and the surface of the window also cools by ~510 C, causing the reference point for Equation 1 to shift.

vacuum port dewar vacuum shroud LN2 entrance and exit ports

dewar thermal shroud and blanket

Cu thermal shroud and blanket test mirror

shutter mechanism dewar window ( 150 mm)

Fizeau interferometer

tip/tilt plate and rail carrier

Figure 3. Cartoon of the null test setup for the flat mirrors. Figure error is measured at 293 K and 80 K using this setup. The dewar thermal shroud is 300 mm in diameter. The Fizeau interferometer is equipped with a transmission flat. For spherical mirrors, the setup is identical, except for the use of a transmission sphere (f/3.3). For spheres, a radius of curvature test is performed using the rail.

(a)

(b)

(c)

(d)

Figure 4. (a) Mirror (spherical prescription) mounted to dewar cold plate. The mirror is simply supported on two pins underneath (not shown) with its back against the cold finger. The mirror is also secured to the plate with Al or Cu tape. Note the two diodes attached to the top of the mirror substrate. (b) Mirror surrounded by Cu radiation shield (bolted to cold plate) with surrounding thermal blanket. (c) Face-on view of mirror in partially assembled dewar with inner thermal shroud and shutter mechanism housing (Al annulus in front). (d) Fully assembled dewar (vacuum shroud installed) in front of interferometer on vibration-isolated table with vacuum pump attached.

3.2 Dewar window effects: Trouble-shooting and modeling

Halfway through a complete thermal cycle, the mirror reaches its equilibrium temperature of ~80 K and the Al vacuum shroud (the outer skin of the dewar) remains at essentially room temperature. The dewar window is mounted to this shroud via a clamp about its perimeter. The temperature of the perimeter of the window is ~293 K, but the central portion of the window is exposed to a large cold solid angle interior to the dewar (mirror and cold finger). In balance, heat is radiated from the window to the cold interior and the center of the window drops 510 C below room temperature. This radial thermal gradient causes the fused silica window to deform --- the center becomes thinner than the perimeter --- which adds defocus wavefront error to the measurement of mirror figure error. Other workers report this or a similar effect.3639 We address this problem as follows: Since the shutter is bolted to the internal Al radiation shroud and the thermal path to the blades is somewhat circuitous, the temperature of the shutter is ~250 K when the mirror has reached equilibrium at 80 K. This is sufficiently warm to shield the window from the cold solid angle of the mirror when the shutter is closed. This essentially eliminates the radial thermal gradient within ~1 hour and renders the effect of the window on the measurement negligible. Rather, the effect of the window on the system wavefront at 80 K is the same as its effect at 293 K. A significant amount of effort went into trouble-shooting this effect. When we cycled the first flat mirror and measured a large change in power from the warm to cold condition, we considered the possibility of a thermal gradient through the thickness of the mirror, perhaps caused by exposure to a warm solid angle (radiation) or poor thermal contact to the cold finger. Thus, we attached diodes to the sides of the mirror to measure a gradient through the mirror thickness and installed a small radiation shield just around the mirror (Figure 4a and b, respectively). To add to our confusion, a finite element analysis of a thermal gradient of 0.1 C acting through the thickness of the mirror produced a figure error map similar in amplitude and sign to the observed power. A gradient of 0.1 C is less than the smallest gradient that we could reliably measure using diodes on the mirror. However, a thermal model of the entire dewar showed that it was impossible to create such large gradients in the test mirror. Also, measurements of the temperature of the perimeter and center regions of the window and measurements of the window surface figure were sensitive to the shutters state. These observations indicated that a radial thermal gradient in the window is to blame for the observed power. In addition, we used two simple optical models to support our conclusion that the window is to blame for the excess power initially observed in mirrors at 80 K. We used ZemaxTM to ray-trace the effect.* We represented the window surface figure error data as Zernike phase additions to the modeled window surfaces. The model accurately predicted the observed power at 80 K. We also modeled the effect using custom routines written in the Interactive Data Language (IDL). We processed the arrays entry-wise using

mreal = mraw +

(n 1) ( w1 w2) n

(2)

where n is the index of refraction of the fused silica window (scalar), w1 and w2 are the measured surface errors from the front and rear of the window, respectively (arrays), mraw is the measured figure error of the window (array), and mreal is the corrected figure error array (array).40 For this analysis, we assumed that n does not change as a function of temperature and is constant across the window. This calculation was in good agreement with the ray-trace model. The data presented in this paper are not corrected for the deleterious effect of the window. The above modeling indicates that, when the shutter is employed, any residual impact is small compared to the observed changes in figure error from 293 to 80 K (Section 3.4).
3.3. Data reduction A set of raw intensity data from the interferometer is converted to a phase map and a measurement of figure error by Zygo Corp.s MetroProTM software. After figure error data is collected at 293 and 80 K, the first step in the data reduction pipeline
*

Focus Software, Inc. P.O. Gox 18228, Tucson, AZ 85731-8228, USA; Tel.: 520-733-0130 Research Systems, Inc., 4990 Pearl East Circle, Boulder, CO 80301, USA; Tel.: 303-786-9900

is to register the arrays and subtract the 293 K from the 80 K array entry-wise (point-to-point). The resulting array is the change from 293 to 80 K. We then perform a least squares fit of piston and tilt over the full aperture and subtract those terms from the data. Next, we perform a linear regression fit of the first 15 Zernike polynomials (to n=4, m=4) over the largest possible circular sub-aperture.41 As a check, this data processing is carried out using both MetroProTM and custom routines written in IDL. The two techniques are in excellent agreement. For the spherical mirrors, we measured the power via a MetroProTM least squares fit of the equation

z ( x, y ) = C 0 + C1 x + C 2 y + C 3( x 2 + y 2 )

(3)

to data over the full mirror aperture, where z is the amplitude of the figure error, x and y describe position in the mirror aperture, and Cn are coefficients determined by the fit.42 The power value given in Table 2 is the C3 coefficient in units of waves at =632.8 nm. By definition, this term is a factor of two larger than the Zernike power coefficient (n=2, m=1).
3.4. Results The figure error change from 293 to 80 K for the flat and spherical mirrors is shown in Figures 5 and 6, respectively. Statistics associated with the images are given in Tables 1 and 2, respectively. In these figures, the mirror blank grain direction runs left right (horizontal on the page).

Qualitatively, the dominant component of the change in figure error for all of the mirrors is astigmatism, oriented roughly left right. When the amplitude of the change is lower, trefoil is often a dominant component. When the mirror is rotated in its mount about the normal to the optical surface, this change rotates with the mirror (i.e., the figure error change is not related to the mount or setup). Quantitatively, the measured change in figure error for mirrors SR 46 is lower in amplitude (i.e., lower RMS) compared to SR 13, although this trend is not as strong for the spherical mirrors. In addition, the change for SR 46 is somewhat more consistent from thermal cycle to cycle (i.e., the mirrors are a little better behaved). When data from both the flat and spherical mirrors are considered, the SR 5 figure error change is the lowest in amplitude and most consistent. The SR 4 spherical mirror was not properly focused in the test setup (operator error). So the RMS values given in Table 2 were calculated after first subtracting the power term derived from a Zernike fit via linear regression. Therefore, only a circular sub-aperture is displayed in Figure 6. The large amplitude of the power could be dominating the signal from more subtle changes in mirror figure. This is a potential source of systematic error for the SR 4 data set. The rail data for the spherical mirrors was reduced using Equation 1. The change in radius from 293 to 80 K for each mirror and thermal cycle is listed in Table 2. The IRMOS optical design assumes that the total change in radius from 293 K to 80 K is

R = 0.398% R

(4)

where R is mirror radius (i.e., the effective coefficient of thermal expansion is 18.3510-6).43 The statistical and systematic errors listed in Section 3.1.2 make these radius measurements of little more use than a rough check. Systematic errors dominate the standard deviation of the measured change from thermal cycle to cycle. The uncertainties quoted in Table 2 for the change in radius for each thermal cycle are an estimate of the measurement error.
3.5. Microroughness Testing We used an ADE Phase Shift MicroXAM white light interferometer to measure the microroughness at ~5 locations on the aperture of each test mirror.* The interferometer has a spatial resolution of 0.3 nm and an amplitude sensitivity of 0.01 nm over a 0.30 0.41 mm field of view. Its software uses a fringe-fitting algorithm to calculate surface error. For powered mirrors, low order, spherical and cylindrical terms are removed by subtraction of a least squares fit to the surface error array.

ADE Phase Shift, 3470 East Universal Way, Tucson, AZ 85706, USA; Tel.: 520-573-9250

The microroughness of the SR 16 plate stock test mirrors is virtually identical in amplitude ( 10 nm RMS) and character. The data appear to be dominated by the interaction of the SPDT tool with the impurities associated with Al 6061. The grain structure and tool marks give the roughness a well-defined orientation for various spatial periods. The microroughness of the mirrors cut from forging stock is about the same amplitude, but much more random in character (Section 5.2).

TC 1

TC 2

TC 3

TC 1

TC 2

TC 3

0.206 RMS

0.195 RMS

0.160 RMS

0.065 RMS

0.067 RMS

0.060 RMS

0.149 RMS

0.142 RMS

0.135 RMS

0.077 RMS

0.064 RMS

0.056 RMS

0.129 RMS

0.140 RMS (a)

0.160 RMS

0.095 RMS

0.076 RMS (b)

0.078 RMS

Figure 5. For flat mirrors, grayscale images of the array representing the change in figure error from 293 to 80 K for a circular sub-aperture as a function of stress relief (SR) process and thermal cycle (TC). Data for SR 13 (top-to-bottom) are shown in column (a) for three consecutive thermal cycles (left to right, TC13). Data for SR 46 are shown in column (b). The standard deviation (RMS) for each surface is given below each image ( = 632.8 nm). The grayscale is not consistent from image to image. Lower areas are darker and higher areas are lighter. The grain orientation is left right.

PROCESS 1

PARAMETER

TC 1

TC 2

TC 3

AVE STDEV

0.206 0.195 0.160 0.187 RMS () 1.145 1.054 0.983 1.061 P-V () Z-FOCUS 0.0151 -0.0444 -0.0754 Z-ASTIG +/-45 0.3755 0.3805 -0.3034 Z-ASTIG 0, 90 -0.3139 -0.2551 0.1978 RMS () P-V () Z-FOCUS Z-ASTIG +/-45 Z-ASTIG 0, 90 RMS () P-V () Z-FOCUS Z-ASTIG +/-45 Z-ASTIG 0, 90 RMS () P-V () Z-FOCUS Z-ASTIG +/-45 Z-ASTIG 0, 90 RMS () P-V () Z-FOCUS Z-ASTIG +/-45 Z-ASTIG 0, 90 RMS () P-V () Z-FOCUS Z-ASTIG +/-45 Z-ASTIG 0, 90 0.149 0.804 -0.0461 0.0244 -0.3482 0.129 0.673 -0.0874 0.2648 -0.1058 0.065 0.449 -0.0885 0.0207 -0.0128 0.077 0.432 -0.0724 0.1308 -0.0327 0.095 0.666 -0.0523 0.1222 -0.1362 0.142 0.807 -0.0456 -0.0157 -0.3313 0.140 0.742 0.0089 0.2913 -0.1709 0.067 0.587 -0.039 0.0666 -0.0798 0.064 0.354 -0.0715 0.0761 0.0204 0.076 0.539 -0.0365 0.0876 -0.1059 0.135 0.142 0.722 0.778 -0.0515 0.0323 -0.3109 0.16 0.143 0.857 0.757 -0.048 0.3362 -0.1541 0.060 0.064 0.395 0.477 -0.0535 0.0897 -0.0439 0.056 0.066 0.401 0.396 -0.0267 0.0874 -0.0396 0.078 0.083 0.439 0.548 -0.0617 0.0814 -0.1103

0.024 0.081

0.007 0.048

0.016 0.093

0.004 0.099

0.011 0.039

0.010 0.114

Table 2. Flat mirror statistics (refer to Figure 5). Only the maximum diameter circular sub-aperture is considered for calculation of standard deviation (RMS) and peak-to-valley (P-V), because the interferometer test leg is only 100 mm in diameter. The magnitude of three dominant Zernike terms is given ( = 632.8 nm).

TC 1

TC 2

TC 3

TC 1

TC 2

TC 3

0.130 RMS

0.121 RMS

0.128 RMS

0.125 RMS

0.126 RMS

0.109 RMS

0.146 RMS

0.137 RMS

0.113 RMS

0.073 RMS

0.058 RMS

0.063 RMS

0.084 RMS

0.094 RMS (a)

0.097 RMS

0.109 RMS

0.087 RMS (b)

0.071 RMS

Figure 6. For spherical mirrors, grayscale images of the array representing the change in figure error from 293 to 80 K for a circular sub-aperture as a function of stress relief (SR) process and thermal cycle (TC). Data for SR 13 (top-to-bottom) are shown in column (a) for three consecutive thermal cycles (left to right, TC13). Data for SR 46 are shown in column (b). The standard deviation (RMS) for each surface is given below each image ( = 632.8 nm). The grayscale is not consistent from image to image. Lower areas are darker and higher areas are lighter. The grain orientation is left right. For SR 4, the mirror was not properly focused in the test setup, and the RMS was calculated after first subtracting the power from a Zernike fit. Only a circular sub-aperture is displayed, because the Zernike polynomials are orthogonal over a circular aperture.

PROCESS PARAMETER 1

TC 1

TC 2

TC 3

AVE STDEV 0.126 0.685

RMS () P-V () Z-ASTIG 45 Z-ASTIG 0, 90 POWER (C3R )


R (%)
2

0.130 0.697 0.1657 -0.1903 -0.063 0.146 0.900 0.1932 -0.1917 -0.062 0.084 0.564 0.1028 -0.1375 -0.012 0.125 0.714 0.1569 -0.2486 -0.751 0.073 0.483 0.0855 -0.0679 -0.035 0.109 0.838 0.1603 -0.049 -0.007

0.121 0.712 0.1551 -0.1697 -0.072 0.137 1.054 0.1937 -0.1496 -0.022 0.094 0.653 0.1022 -0.1618 -0.019 0.126 0.724 0.1306 -0.2533 2.011 0.058 0.443 0.0524 -0.0507 -0.032 0.087 0.601 0.1217 -0.0184 0.016

0.128 0.645 0.1645 -0.1509 -0.234 0.113 0.709 0.1447 -0.1421 -0.042 0.097 0.669 0.1316 -0.1255 -0.09 0.109 0.670 0.148 -0.2072 -0.338 0.063 0.540 0.0612 -0.0368 0.086 0.071 0.594 0.0859 -0.0401 -0.019

0.005 0.035

-0.430.01 -0.410.01 -0.460.01 -0.43


0.132 0.888

0.02 0.017 0.173

RMS () P-V () Z-ASTIG 45 Z-ASTIG 0, 90 POWER (C3R )


R (%)
2

-0.380.01 -0.400.02 -0.410.01 -0.40


0.092 0.629

0.02 0.007 0.057

RMS () P-V () Z-ASTIG 45 Z-ASTIG 0, 90 POWER (C3R )


R (%)
2

MISSING -0.330.01 -0.390.01 -0.36


0.120 0.703

0.04 0.010 0.029

RMS () P-V () Z-ASTIG 45 Z-ASTIG 0, 90 POWER (C3R )


R (%)
2

-0.270.01 -0.610.02 -0.390.01 -0.42


0.065 0.489

0.17 0.008 0.049

RMS () P-V () Z-ASTIG 45 Z-ASTIG 0, 90 POWER (C3R )


R (%)
2

-0.330.01 MISSING -0.540.01 -0.43


0.089 0.678

0.15 0.019 0.139

RMS () P-V () Z-ASTIG 45 Z-ASTIG 0, 90 POWER (C3R )


R (%)
2

-0.370.01 -0.400.01 -0.380.01 -0.38

0.02

Table 3. Spherical mirror statistics (refer to Figure 6). The full, rectangular mirror aperture is considered for calculation of standard deviation (RMS) and peak-to-valley (P-V), except for SR 4, where the maximum diameter circular sub-aperture was used. The magnitude of two dominant Zernike terms is given for the maximum diameter circular sub-aperture ( = 632.8 nm). The power value is discussed in Section 3.3 (Equation 3).

4. STRENGH AND MATERIALS TESTING


4.1. Strength and strain testing We report the results of tensile testing to determine the yield strength, ultimate strength, elongation, and proportional limit of samples from SR 16 (Section 2.1). Dog bone tensile samples are removed from the plate stock along with each test mirror and follow that mirror through its SR process (Figure 1b). Six tensile samples are fabricated for each SR process. They are all oriented long transverse in the plate and are 200 mm long with a gage length cross section of approximately 6 13 mm.44

Figure 7. Dogbone tensile sample in Instron 1125/4400R Universal test machine using a 89,000 N load cell and MTS hydraulic wedge grips.

Testing is performed on an Instron 1125/4400R Universal Test Machine using a 89,000 N load cell with model number A30-33 and serial number 295. The specimens are held using MTS hydraulic wedge grips with model number 647 (Figure 7). An averaging extensometer with model number 231-1002 and serial number 7279 is used to measure strain to 5%. Two specimens from each SR are loaded to failure using a crosshead speed of 1.27 mm/min. A third specimen from each SR is loaded and unloaded to increasingly higher strains. Its permanent deformation is measured each time the specimen was unloaded.

4.2. Strength results

A total of thirty-six tensile specimens are tested. The individual test results are presented elesewhere.31 Loading curves are shown in Figure 8. Loading curves showing permanent deformation as a function of applied stress are shown in Figure 9. A representative loading curve that shows multiple loadings and un-loadings, used to measure permanent deformation between 0 and 400 microstrain, is shown in Figure 10. The uncertainty associated with our extensometer and the axiality of our pull test is probably not adequate to measure mechanical properties at 10 microstrain plastic deformation. However, the trends measured at 50 microstrain are less uncertain and consistent with those at 10 microstrain.
7000 6000 5000 4000 Applied Load 3000 (lbs) 2000 1000 0
Figure 8. Representative loading curves. Cond. n corresponds to SR n in the text.

Cond. 4 Cond. 5 & 6 Cond. 1

Cond. 2 & 3

0.05

0.1

0.15

0.2

0.25

0.3

Crosshead Movement (in)

40000 35000 30000 25000 20000 15000 10000 5000 Condition 1 Condition 2 Condition 3 Condition 4 Condition 5 Condition 6 0 50 100 150 200 250 300 350 400

Tensile Stress (psi)

Figure 9. Loading curves showing the maximum applied tensile stress and the associated permanent deformation measured in units of strain. Condition n corresponds to SR n in the text.

Permanent Deformation ( strain)

3000

003A-3
2500 2000 Strain 1500 1000 500 0
Figure 10. Loading and unloading to increasingly higher strains to measure permanent deformation for a sample associated with flat mirror SR 3.

200

400

600

800

1000

Time (sec)

Permanent deformation

Samples from SR 2 and 3 (both overaged at 260 C for 2 hours) show markedly reduced macro and micro strengths, while completely re-heat treating (SR 46) results in slightly elevated mechanical properties compared to the as-received condition (SR 1). There are no significant differences in mechanical properties between the samples removed near the center of the original plate and those removed near the plate edge, nor between samples that compared the effects of uphill quenching (SR 2 vs. 3 and 4 vs. 5). The ultimate and yield strengths of the glycol-quenched samples, SR4 and SR5, are slightly higher than for the water-quenched samples, SR6. The proportional limit values at 10 microstrain plastic strain are of limited accuracy because they are derived from only one sample per condition using a relatively low resolution strain measuring instrument (i.e., an extensometer). The data illustrate the deleterious effect of a 260 C stress relief and the beneficial effects of completely re-heat treating (Table 3).

Stress Relief Conditioning Requirement SR 1 SR 2 SR 3 SR 4 SR 5 SR 6

Ultimate Yield Strength (ksi)


42.0 42.9 33.3 33.3 46.3 46.2 44.7

Yield Strength (ksi)


35.0 38.1 27.3 27.8 41.6 41.3 40.0

Elongation (%)
9.0 15.4 15.3 15.2 16.4 16.6 16.1

Proportional Limit* (ksi)


N/A 22.1 12.8 16.9 33.3 31.0 31.0

Table 3. Minimum requirements, average test results, and proportional limit for stress relieved Al 6061-T651. 4.2. Grain structure Metallographic cross sections corresponding to the polished mirror surfaces are shown in Figure 11 for the 25 mm plate stock (Figures 11a and 11b) and for the forgings (Figures 11c and 11d). The as-polished plate stock (Figure 11a) shows an insoluble second phase particle network that typically consists of Fe-Mn-Cr-Si-Al compounds. The as-polished forging (Figure 11c) also contains second phase particles, but they are more uniformly distributed and of a finer size. The microroughness of mirrors made from plate and forgings was about the same (Section 3.5). Corresponding structures after etching with Kellers process show, for the plate (Figure 11b), an elongated grain structure from rolling and some color contrast indicative of chemical or precipitation inhomogeneity, and, for the forging (Figure 11d), a finer grained, more equiaxed and chemically uniform structure. The plate is processed via SR 5 and the forging via SR 4.

This is the stress at which 10 microstrain of permanent deformation was measured. These values are derived from only one sample per SR using an extensometer, but they show the deleterious effect of a 260 C stress relief and the beneficial effect of completely reheat treating.

(a)

(b)

(c)

(d)

Figure 11. (a) and (b): 2.5 2.5 mm images of 25 mm plate, processed per SR 5, with orientation corresponding to the polished mirror surfaces (longitudinal/long transverse) . (a) As-polished. (b) Etched via Kellers process. (c) and (d): 2.5 2.5 mm images of a forging, processed per SR 4, with orientation corresponding to the polished mirror surfaces (probably transverse) . (c) Aspolished. (d) Etched via Kellers process.

5. CONCLUSION
SR 2 and 3 show reduced macro and micro strengths, while SR 46 (i.e., complete re-heat treating) show the same or elevated mechanical properties compared to as-delivered Al 6061-T651. There is a clear improvement in mirror figure error change for SR 46 when compared to SR 13. This establishes that a high glycol content quench with or without uphill quenching and a room temperature water quench plus uphill quench can result in lower residual stress than mechanically stress relieved material even with a significant overage stress relief. Systematic errors in the figure data aside, the SR5 mirrors show improved dimensional stability over SR 4. This indicates that the addition of the uphill quench to SR 4 further reduces residual stress from the glycol (downhill) quench. The uphill quenching is performed immediately after the glycol quench on a part that has a uniform residual stress pattern and minimum strength and maximum plasticity. We select SR 5 for the IRMOS mirrors, because the SR 5 test mirror figure error change from 293 K to 80 K is generally lower in amplitude than the other processes and more consistent over multiple thermal cycles.45 The presence of large systematic uncertainties make the radius data inconclusive, but the data are consistent with Equation 4. The SR process that we implement for the IRMOS instrument structural components (i.e., optical bench) is simpler than that used for the mirrors, because the dimensional stability requirements are not as demanding.46 The process involves a stress relief heat treatment and a series of thermal cycles between rough and finish machining fabrication steps.

ACKNOWLEDGEMENTS
We are indebted to Roger A. Paquin for his help in planning this experiment. For technical support and useful discussions, we gratefully acknowledge: Dr. Charles Bowers, Dr. David A. Content, William L. Eichhorn, Sandra M. Irish, Mark B. Mann, and Armando Morell of GSFC; Russell B. Makidon and Robert S. Winsor of STScI; Thomas J. French and Peter Petrone of ManTech International Corp.; Christopher May of Orbital Science Corp.; Shelly Conkey and Rita Rosenbaum of Swales Aerospace, Inc.; Steven J. Conard of the Johns Hopkins University Instrument Development Group; and Alex Sohn of the North Carolina State University Precision Engineering Center. This work is supported by the Next Generation Space Telescope project at GSFC.

REFERENCES
1. 2. 3. R. S. Winsor, J. W. MacKenty, M. Stiavelli, M. A. Greenhouse, J. E. Mentzell, R. G. Ohl, and R. F. Green, Optical design for an infrared imaging multi-object spectrometer (IRMOS), Proc. SPIE 4092, pp. 102108, 2000. J. Zimmerman, Strain free mounting techniques for metal mirrors, Optical Engineering 20, pp. 187189, 1981. R. G. Ohl, W. Preuss, A. Sohn, S. Conkey, K. Garrard, J. G. Hagopian, J. M. Howard, J. E. Hylan, S. M. Irish, J. E. Mentzell, M. Schroeder, L. M. Sparr, R. S. Winsor, S. W. Zewari, M. A. Greenhouse, and J. W. MacKenty, Design and fabrication of diamond machined, aspheric mirrors for ground-based, near-IR astronomy, Proc. SPIE 4841, 2002, accepted. C. W. Marschall and R. E. Maringer, Dimensional Instability: An Introduction, Pergamon Press, New York, 1977. C. W. Marschall and H. E. Hagy, Dimensional stability workshop, Proc. SPIE 1335, pp. 217243, 1990. R. A. Paquin, Dimensional instability of materials: how critical is it in the design of optical instruments? Proc. SPIE CR43, pp. 160180, 1992. R. A. Paquin, Metal Mirrors, Handbook of Optomechanical Engineering, A. Ahmad, ed., CRC Press LLC, New York, pp. 89110, 1997. Alloy Temper Designation Systems for Aluminum and Aluminum Alloys, ASM Handbook, Volume 2, Properties and Selection: Nonferrous Alloys and Special Purpose Materials, ASM International, Materials Park, pp. 1628, 1990. Aluminum Mill and Engineered Wrought Products, ASM Handbook, Volume 2, Properties and Selection: Nonferrous Alloys and Special Purpose Materials, ASM International, Materials Park, pp. 2961, 1990. Heat Treating of Aluminum Alloys, ASM Handbook, Volume 4, Heat Treating, ASM International, Materials Park, pp. 841879, 1991. Heat Treatment of Wrought Aluminum Alloy Parts, Aerospace Material Specification 2770E, Society of Automotive Engineers, Inc., Warrendale, 1989. M. R. Howells and R. A. Paquin, Optical substrate materials for synchrotron radiation beam lines, Proc. SPIE CR67, pp. 339372, 1997. M. C. Gerchman, Specifications and manufacturing considerations of diamond machined optical components, Proc. SPIE 607, pp. 3645, 1986. J. B. C. Fuller, Jr., P. Forney, and C. M. Klug, Deisgn and fabrication of aluminum mirrors for a large aperture precision collimator operating at cryogenic temperatures, Proc. SPIE 288, pp. 104110, 1981. T. J. Magner and R. D. Barney, Interferometric phase measurement of zerodur, aluminum and SXA mirrors at cryogenic temperatures, Proc. SPIE 929, pp. 2938, 1988. A. A. Ogloza, D. L. Decker, P. C. Archibald, D. A. OConnor, and E. R. Bueltmann, Optical properties and thermal stability of single-point diamond-machined aluminum alloys, Proc. SPIE 966, pp. 228251, 1988. S. C. Fawcett and the Cassini/Composite Infrared Spectrometer (CIRS) project, Stress Relieve Heat Treatment Procedure for CIRS Aluminum Mirrors, CIRS PROC-EB53-HT-4, NASA/Marshall Space Flight Center, Huntsville, 14 July 1994. S. C. Fawcett and the Cassini/Composite Infrared Spectrometer (CIRS) project, Fabrication Procedure for CIRS Aluminum Mirror 1532592, CIRS PROC-EB53-592-4, MSFC Drawing 96M79022, NASA/Marshall Space Flight Center, Huntsville, 14 July 1994. S. C. Fawcett, Development of Thermally Stabilized Aluminum Mirrors for the Composite Infrared Spectrometer, Proceedings of ASPE 9th Annual Meeting, pp. 346349, 1994. R. G. Probst, T. Ellis, A. M. Powler, I. Gatley, G. Heim, and K. M. Merrill, Cryogenic optical bench: a multifunction camera for infrared astronomy, Proc. SPIE 2198, pp. 695702, 1994. D. Vukobratovich, K. Don, and R. E. Sumner, Improved cryogenic aluminum mirrors, Proc. SPIE 3435, pp. 918, 1998.

4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21.

22. T. Hegarty and the Space Infrared Telescope Facility/Infrared Array Camera (IRAC) project, Fabrication and Stress Relieve Procedure for IRAC Aluminum Mirrors, IRAC 544-PROC-003, NASA/Goddard Space Flight Center, Greenbelt, 22 February 1999. 23. N. Kobayashi, Subaru Telescope, National Astronomical Observatory of Japan, personal communication, 2000. 24. J. E. Hylan and the Infrared Multi-Object Spectrograph (IRMOS) project, IRMOS drawing number 2045973, NASA/Goddard Space Flight Center, Greenbelt, 2000. 25. J. A. Connelly, R. G. Ohl, T. T. Saha, T. Hadjimichael, J. E. Mentzell, R. G. Mink, V. J. Chambers, M. A. Greenhouse, R. S. Winsor, and J. W. MacKenty, Subsystem imaging performance and modeling of the Infrared Multi-Object Spectrograph, Proc. SPIE 4841, 2002, accepted. 26. R. W. Toland, R. G. Ohl, M. P. Barthelmy, S. W. Zewari, J. C. McMann, and R. G. Mink, Comparison of stress relief procedures for cryogenic aluminum mirrors, Optical Engineering, 2002, in preparation. 27. Alcoa Green Letter: Avoiding Stress-Corrosion Cracking in High Strength Aluminum Alloy Structures, pp. 2023, Alcoa Corp., Pittsburgh, 1982. 28. D. M. Walker and R. Y. Hom, Residual Stress Analysis of Aircraft Aluminum Forgings, Advanced Materials and Processes, June, 2002. 29. R. A. Paquin, Process Recommendations for IRMOS Aluminum Mirror and Structural Components, report for consulting agreement under NASA PO number S-40873-G, NASA/Goddard Space Flight Center, Greenbelt, 2000. 30. R. A. Paquin, Comments and Recommendations on IRMOS Mirror Prefabrication Test Program (IRMOS-301), report for consulting agreement under NASA PO number S-43044-G, NASA/Goddard Space Flight Center, Greenbelt, 2000. 31. M. P. Barthelmy, R. G. Ohl, and S. W. Zewari, Stress relief procedures for cryogenic dimensional stability of aluminum mirrors and structures: Fabrication procedures and test data, NASA technical memo, NASA/Goddard Space Flight Center, Greenbelt, 2003, in preparation. 32. Heat Treating, Metals Handbook, 9th Edition, Volume 4, pp. 696697, American Society of Metals, 1981. 33. Heat Treatment of Wrought Aluminum Alloy Parts, AMS 2770. 34. P. R. Yoder, Opto-Mechanical Systems Design, pp. 353355, Marcel Dekker, Inc., New York, 1992. 35. D. Malacara, Optical Shop Testing, 2nd Edition, pp. 585286, John Wiley and Sons, Inc., New York, 1992. 36. D. Vukobratovich, Optomechanical System Design, Infrared and Electro-Optical Systems Handbook, Volume 4, M. Dudzik, ed., pp. 123126, International Society for Optical Engineering (SPIE), Bellingham, 1993. 37. D.-H. Gwo, Simultaneous interferometric optical figure characterizations for nominal optical flat and dewar-window system at cryogenic temperatures, Proc. SPIE 2814, pp. 3645, 1996. 38. D. Vukobratovich, Optomechancial Design Principles, Handbook of Optomechanical Engineering, A. Ahmad, ed., pp. 5558, CRC Press LLC, New York, 1997. 39. D. R. Hearn, Vacuum window optical power induced by temperature gradients, Proc. SPIE 3750, pp. 297308, 1999. 40. C. Bowers, NASA/Goddard Space Flight Center, personal communication, 2002. 41. D. Malacara, Optical Shop Testing, 2nd Edition, p. 465, John Wiley and Sons, Inc., New York, 1992. 42. MetroProTM Reference Guide, OMP-0347, p. 3-46, Zygo Corporation, Middlefield, 1997. 43. Y. S. Touloukian, Thermophysical Properties of Matter, Thermal ExpansionMetallic Elements and Alloys, IFI / Plenum, New York, 1975. 44. E8 Standard Test Methods for Tension Testing of Metallic Materials, Annual Book of ASTM Standards, American Society for Testing and Materials, West Conshohocken. 45. M. P. Barthelmy and the Infrared Multi-Object Spectrograph (IRMOS) project, IRMOS-303, Heat Treatment and Thermal Stabilization of IRMOS Mirrors and Grating Blanks, NASA/Goddard Space Flight Center, Greenbelt, May 2001. 46. M. P. Barthelmy and the Infrared Multi-Object Spectrograph (IRMOS) project, IRMOS-330, Stress Relief and Thermal Cycling of IRMOS Structural Components, NASA/Goddard Space Flight Center, Greenbelt, April 2002.

Vous aimerez peut-être aussi