Chapter 1 Introduction

The performance of high-precision optical systems using spherical optics is limited by aberrations. By applying aspherical and freeform optics, the geometrical aberrations can be reduced or eliminated while at the same time also reducing the required number of components, the size and the weight of the system. New manufacturing techniques enable creation of high-precision freeform surfaces. Suitable metrology is key in the manufacturing and application of these surfaces, but not yet available. Objective of the NANOMEFOS project is to realize a new metrology method to complete the freeform surface manufacturing value chain.

1.1 Aspherical and freeform optics

High-precision optical systems can for instance be found in space, science and lithography applications. Well known examples are various earth and space observation satellites, lithography projection systems, astronomical telescopes and optical microscopes. Consumer products increasingly also incorporate precision optical systems, such as digital cameras, multi-media projectors and car head-up displays. Miniature optics are found in optical storage and camera phones. 

The requirements for these systems are ever increasing towards better, lighter, cheaper, smaller and also larger. Most optical systems employ flat and spherical optics due to the relative ease of manufacturing and measuring. Their performance is, however, limited by aberrations. Applying aspherical and freeform surfaces allows for elimination of these aberrations, combined with many other advantages. Their application requires new techniques for optical designing, manufacturing and measuring. This thesis describes the development of a new metrology method for aspherical and freeform optical surfaces. 

1.1.1 Advantages of aspherical and freeform optics

Geometrical optical aberrations are inherent to the use of spherical optics. Figure 1.1 (top) schematically shows spherical aberration of a spherical lens. The outer rays are deflected more than the inner rays, giving a distorted focus. Currently, aberrations are mostly reduced by applying multiple spherical elements in series. A more elegant way to reduce aberrations is by applying aspherical and freeform optics, as illustrated by Figure 1.1 (bottom). 

Spherical lens

Aspherical lens

 

Figure 1.1: Elimination of spherical aberration with an asphere

Besides improving the optical quality of a system, application of aspheres and freeforms also allows for a reduction of the required number of components, and with that also a decrease of stray light and of system size and mass (ASPE, 2004; OptoNet 2006; OptoNet 2008). In (Winsor et al., 2004) and (Garrard et al., 2005) it is for instance reported that an order of magnitude size reduction of the ground based infrared IRMOS spectrometer is obtained by utilizing freeform surfaces. They further report reduced alignment efforts due to the fewer number of elements. Dynamic optical motion systems, such as optical storage focusing objectives, also benefit from reduced size and weight, but these optics are generally very small and are not the subject of this thesis.

Freeform optics allow for more design flexibility. Since there are relatively few degrees of freedom in a spherical system (such as the radii of curvature, thickness and diameter, the distance between the elements and the materials used), the boundary conditions for the mechanical design are usually dictated by the optical layout. This is especially the case in for instance satellite instruments, where a maximum functionality is required in a minimum of design volume, combined with harsh launching and environmental conditions. By applying aspherical and freeform optics, the optical layout becomes more flexible, enabling it for instance to depart from rotational symmetry. This allows for a more optimal opto-mechatronic design. In retrofitting it is the other way around; the mechanics dictate the optical design space.

An example is the SCUBA-2 instrument (Atad-Ettedgui et al., 2006), where the optical requirements in combination with the available design volume led to a freeform optical design.

The increased number of degrees of freedom in the surface shape however also has an almost infinite number of solutions with minimal aberrations. The challenge here is now no longer in minimizing the aberrations of the nominal system, but in optimizing the alignment and manufacturing tolerances (Garrard et al., 2005), to truly benefit from the extra design freedom.

Another field where freeform optics have advantages is in illumination optics (Ries and Muschaweck, 2002), such as car headlight reflectors. Here, freeform optics allow for far more effective tuning of the spatial illumination distribution than with aspherical optics.

1.1.2   Applications and trend

The evolution of aspherical optics is described in (Heynacher, 1979). The concept of using aspherical optics was first suggested by Kepler in 1611, and given a theoretical foundation by Descartes in 1638. This has mainly remained theoretical practice until the arrival of computer-controlled manufacturing techniques. The development of CNC machining processes has enabled generation of aspheres with sufficient accuracy for early applications, especially in large volume production and dies for moulded optics. The advantages are now widely recognized, but due to the fabrication and metrology issues, optical designers are still reluctant to apply aspheres and freeforms (Garrard et al., 2005).

Aspheres

Moderate precision aspheres are currently becoming widely adopted in mass production consumer applications, such as (semi-)professional camera objectives. Figure 1.2 (left) for example shows an objective where the application of an asphere led to half the size and weight and increased image quality (Schuhmann, 2007).

 

Figure 1.2: Improvement with aspheres of a camera objective (Schuhmann, 2007) and DUV projection system (Ulrich et al., 2003)

High-precision aspheres are mainly being applied in single-piece and small series diffraction limited applications, such as science and lithography instruments. Figure 1.2 (right) for example shows the size and number of elements reduction that was achieved by using aspheres for a DUV lithography projection system (Ulrich et al., 2003). In the next generation Extreme UV lithography systems aspherical off-axis mirrors are applied for which accuracies of better than 1 nm are required (Dörband and Seitz, 2001). Large high-precision aspheres are increasingly applied in astronomical telescope designs, such as the Hubble space telescope (Allen et al., 1990) and the European Extremely Large Telescope (Gilmozzi and Spiromilio, 2007).

Freeforms

The ophtalmics industry already commonly applies freeform optics in the form of multi-focal and cylinder corrected spectacle glasses. The required accuracy in this case is in the micrometer range, since the human eye is able to compensate the remaining error. Together with illumination optics, these are probably the most common example of freeform optics, but they are not subject in this thesis since the challenge in producing these optics is more in cost and speed, rather than in form accuracy.

 

One of the earliest applications of imaging freeform optics is the Polaroid SX-70 folding Single Lens Reflex camera (Figure 1.3), introduced in 1972 (Plummer, 1982). Here, the foldable design and the off-axis viewing optical system required two freeform lenses to provide a well corrected system. A more recent example where the mechanics dictated the optical layout is the SCUBA-2 instrument (Atad-Ettedgui et al., 2006; Saunders et al., 2005). For this infrared instrument the required accuracy was 3-15 m. In the visible spectrum, moderate precision freeform surfaces are starting to be applied in for instance head-up and helmet mounted displays (Zänkert, 2008), where size and weight are critical. Applications of high-precision (diffraction limited) freeform optics for visible spectrum imaging systems are still rare.

Trend

As observed by (Beckstette, 2008), the trend for freeform optics is repeating the history of aspherical optics. The first applications are now found in low accuracy high volume (moulded) optics in reflectors and spectacle glasses, and high-end specialties such as spectrometers and off-axis telescopes. The same drivers that stimulated the development of aspheres (size, mass, number of components etc.) now continue to do the same freeforms. The development of new technical capabilities in design, manufacturing and metrology will lead to more high precision single piece and small series applications. With aspheres, the paradigm has recently changed from “Only use aspheres if you have no other way” to “Keep in mind aspheres can help! Learn how to use them”. According to (Beckstette, 2008), comparison of conventional and freeform design examples shows that freeforms are “The winner under all aspects”. Freeform optics are therefore expected to eventually become a commodity just as aspheres.

 

Further supporting the observed trend is the increasing number of scientific conferences and publications dedicated to all aspects of aspherical and freeform optics, by among others ASPE (ASPE, 2004), OptoNet (OptoNet, 2006; OptoNet 2008), Euspen (Euspen, 2009), and various SPIE conferences. Large companies such as Carl Zeiss (Beckstette, 2008), EADS Astrium (Holota, 2008), Berliner Glass (Schuhmann, 2007), ASML (ASML, 2004) also increasingly recognize the potential of aspherical and freeform optics. The Freeform Fabrication & Metrology program of TNO (Saunders et al., 2005) is also an example of this development.

Metrology

Application of high-precision aspherical and freeform optics requires new technologies to be developed in all fields of opto-mechanical system development: in the area of optical design, fabrication, metrology, alignment and assembly. Manufacturing technologies are available now (see section 1.1.4), but metrology is currently believed to be a lacking key enabling technology. Many fabricators at the conferences mentioned before, find that there is no suitable metrology method for aspherical and freeform surfaces, especially for surfaces with large departures from spherical. Citing the recent (NIST, 2008): “Measuring aspheric surfaces poses formidable metrology problems because of the difficulty of obtaining a reference wavefront that closely matches the desired form of the asphere. No single, widely recognized, general, validated way exists for calibrating or measuring complex surfaces with nm-level uncertainties”.

Suitable metrology will enable generation of the optics to the required accuracy, which in turn will encourage optical designers to incorporate them into new designs. This will generate expertise and trust, and lowers manufacturing cost, towards the more common application of aspheres and freeforms.

1.1.3 Surface properties and definitions

Many types of optical surfaces exist, with many different describing terminologies. In (ISO, 2006), standards for spherical optics are described extensively, but aspherical optics are only briefly addressed. No unified terminology is yet defined for freeforms. This section defines the surface properties and terminology as used in this thesis.

Surface types

In Figure 1.4 the different surface shape types are defined. The departure from spherical is exaggerated for explanatory purposes. The term asphere is used by some for all non-spherical optics, leading to confusion on rotational symmetry. In this thesis aspherical surfaces are therefore defined as rotationally symmetric but non-spherical. Examples are parabolic, elliptical and higher order rotationally symmetric surfaces. 

A freeform surface is non-rotationally symmetric and can thus be virtually any shape. Generally, they can be approximated by smooth waves superimposed on an aspherical best-fit. Surface form is usually expressed in x,y,z coordinates, a polynomial description or Zernike coefficients. Flats, spheres and aspheres can be considered limit cases of a freeform surface; the ability to measure freeforms also enables measurement of rotationally symmetric surfaces.

 

                                   Asphere                                 Freeform

 

                                   Off-axis                           Off-axis freeform

 

Figure 1.4: Aspherical, freeform and off-axis surface examples (departure from spherical exaggerated for explanatory purposes)

An off-axis surface is a part of a larger surface that fits one of the previous surface types. An off-axis surface can be considered to be rotationally symmetric or freeform, depending on the chosen position of the axis of rotation.

The circumference of any of the surface types can be non-circular, such as square or elliptical. There can be holes in the surface, for instance for light to pass into or out of the system. For illustration purposes, an extreme example of an off-axis freeform with an off-axis hole is shown in Figure 1.4, together with the original on-axis surface it was a part of.

Dimensions

Optics manufacturing machines are available roughly up to ∅200 mm, up to ∅500 mm (www14, 15, 16 and 22) and custom machines for diameters as large as 8.4 meters (Martin et al., 2003). 

Departure from spherical

Currently, a departure from the best-fit-sphere of several tens of micrometers results in an unresolvable amount of fringes on a Phase-Shifting Interferometer (section 1.1.5 and Appendix B) and is therefore currently considered large. It is suspected that the lack of a metrology method is restricting the departure currently applied by optical designers, so larger departures are expected in future applications. The aspherical departure is chosen to be virtually unlimited, and the Peak-to-Valley freeform departure is chosen to be limited to 5 mm. These large departures may especially occur with infrared optics and off-axis optics that are measured on-axis.

It is expected that approximately 80% of the surfaces will have less than a few millimeters aspherical, and less than a few tenths of a millimeter freeform departure. For illustrational purposes: (Chang et al., 1997) surveyed thousands of lens designs and patents, showing that about 85% has an aspherical departure of the best-fit-sphere of less than 3000 waves, or about 1.8 mm in the visible spectrum (λ =600 nm). 

 

Figure 1.5: Aspherical departure from best-fit-sphere (Chang et al., 1997)

Global and local surface slope

Surfaces can be convex as well as concave. The global slope can range up to 90° for a hemispherical or cylindrical surface, but generally amounts up to 30°. The waves superimposed on the aspherical best-fit, cause local slope variations on top of the global slope. Depending on the typical wavelength and amplitude, unidirectional local slope variations up to 5° are expected. 

This typically means that that a sine with 20 mm wavelength can have 0.55 mm PV, and that at the outer edge of a ∅500 mm part almost 9 waves with 5 mm PV are allowed. Most optics, however, will have less local slope, in the order of a few tenths of a degree. 

Local radius of curvature

Optical surfaces as meant in this thesis are generally very smoothly curved. Besides the holes there are no discontinuities such as steps present. The global radius of curvature (Rc) is usually limited to several tens of millimeters for small lenses, as is the local radius of curvature for high order surface form components.

Materials and reflectivity

Transmission optics can be made from a large variety of glass and other transparent ceramics. Reflectivity of the uncoated optics is generally around 4% for visible light. Reflective optics can either be a metal-coated glass or ceramic substrate, or a mirror machined directly onto a metallic substrate. Reflectivity then ranges from 60 to (theoretically) 100%. Form measurements on surfaces during production are performed on surfaces without a (multi-layer) anti-reflection or protective coating. 

Form, waviness and roughness

According to (Whitehouse, 1994), surface topography can be split in the spatial frequency domain into form, waviness and roughness. Form refers to the macroscopic shape, and for optics covers spatial wavelengths down to about 1 mm. Waviness, caused by for instance tool marks and instabilities in the machining process is in the range of approximately 1 mm down to 20 μm. Roughness is covered by the spatial frequency range below 20 μm. 

Expressing the surface topography in a PSD spectrum is increasingly applied to characterize the surface (Duparré et al., 2002; Youngworth et al., 2005). For high-end freeform surfaces, the content of the form range is in the order of millimeters, while the content of the waviness and roughness range decreases from micrometers to nanometers during production. Especially waviness (mid-spatials) is undesirable in an optical surface. 

The required form accuracy for moderate precision optics is around λ/4, where λ/20 is required for high precision diffraction limited optics. For visible light this correlates to 150 nm and 30 nm, respectively. DUV and EUV optics require an even higher accuracy (λ/100). These are beyond the scope of this thesis.

Scratch / dig

A commonly used method to specify surface defects is by a scratch/dig number (MILPRF-13830B). This number quantifies the allowable amount of scratches and digs in a surface, and has a more aesthetic meaning rather than functional. Scratches typically have a width in the order of tens of micrometers, and pits have a typical diameter in the order of tenths of millimeters. The scratch/dig number is usually determined visually.

1.1.4   Current manufacturing methods

Classical polishing methods used for spherical surfaces (see also Appendix A), are not applicable to general aspheres and freeforms (Karow, 1993). New computercontrolled local polishing methods enable the generation of precision complex surfaces. Examples are the Zeeko Precessions process (Walker et al., 2001), the QED MRF process (Dumas et al., 2004), and ion- and plasma beam machining (Frost et al., 2003). These processes have a deterministic influence function, allowing for rapid iteration in corrective polishing to the desired form. 

For non-ferro metals and some ceramics, single point diamond turning (SPDT) is applied for generating aspheres. Non-rotationally symmetric surfaces can be made by SPDT with a slow- or fast-tool-servo or by precision grinding (Wanders, 2006). Diamond turning or precision grinding is also applied to manufacture moulds for glass pressing or hot embossing techniques, with which large series of complex optics can be manufactured cost effectively (Allen et al., 2006).

1.1.5   Current metrology methods

In every fabrication process the achievable precision is only as good as the measurement method. For a high-end, single piece, freeform optics production environment, important characteristics of a measurement method are: • high accuracy

  • universal
  • non-contact
  • large measurement volume
  • short measurement time

 

Many techniques, experimental as well as commercial, exist for measuring optics. A recent overview of measuring general freeform surfaces, ranging from car body parts to turbine blades to optics, is given by (Savio et al., 2007). In Appendix B the existing methods are assessed with respect to the above characteristics and the properties defined in section 1.1.3. A brief summary is given here.

 

The metrology methods can be separated in imaging and scanning techniques. The main imaging technique is Phase-Shifting Interferometry and the main scanning techniques are stylus profilers and coordinate measuring machines. Some experimental methods are also discussed.

Interferometry

Phase-Shifting Interferometry (Figure 1.6) currently is the work horse for measuring flat and spherical optics. An extensive overview of many interferometry based techniques is given in (Malacara, 2007). The entire surface is imaged at once, measuring it in seconds. With a calibrated reference surface and proper measurement conditions, uncertainty can be of (sub)nanometer order (Dörband and Seitz, 2001). Apertures are generally around 100 mm, but some larger setups exist. 

 

Figure 1.6: Phase-Shifting Interferometer (www26)

If the departure from spherical of the surface under test exceeds several micrometers, the fringes become too dense to resolve, limiting the applicability for aspheres and freeforms. Applying null optics or computer generated holograms (Burge and Wyant, 2004), resolves this issue but at the cost of traceability and universality. Stitching subapertures is sometimes done (Dumas et al., 2004), but large departures require many sub-apertures which gives the risk of stitching errors.

Stylus profilometers

Scanning methods generally provide a more universal measurement compared to imaging techniques. A stylus profilometer (Figure 1.7) performs a 2D line scan with a diamond or ruby tipped stylus (Whitehouse, 1994). This is currently the most commonly applied method for measuring aspheres. For measuring freeforms, a transverse stage is added. In 2D mode, the uncertainty can be in the order of several tens of nanometers. 

 

Figure 1.7: Stylus profilometer (www20)

The measurement length is limited to about 200 mm, and the allowed sag is in the order of 20 mm (www20). The contact stylus requires slow scanning speeds (~10 mm/s) and has the risk of damaging the surface. Transverse inclined surfaces further give rise to torsional deflection errors in the stylus arm and bending of the stylus itself. 

Coordinate Measuring Machines

Truly universal 3D measurements are done with Coordinate Measuring Machines (CMM). Conventional portal CMMs are limited by Abbe errors (Abbe, 1890). Intermediate bodies can be introduced in the guidance setup to keep the linear scales aligned to the probe tip and thus eliminate the Abbe offset in the horizontal plane (Vermeulen, M., 1999; Van Seggelen, 2007). 

In (Ruijl, 2001; Jäger et al., 2001) laser interferometers are used that are orthogonally aligned to a stationary probe tip, and are measuring to a moving mirror block on which the product is mounted. In (Becker and Heynacher, 1987) fixed reference mirrors are applied and the laser interferometers emanate from the moving probe.

The measurement volume of these CMMs is generally less than 1 dm3, except (Becker and Heynacher, 1987), and uncertainties in the order of several tens of nanometers are reported. Contact trigger probes (Weckenmann et al., 2006) are employed. The orthogonal setup makes application of an optical non-contact probe difficult for surfaces with slopes of more than a few degrees.

 

Figure 1.8: CMM with AFM probe (Takeuchi et al., 2004)

In (Takeuchi et al., 2004) an Atomic Force Probe is used (Figure 1.8), in combination with interferometers and stationary reference mirrors. Although the surface is not contacted, scanning speeds are still limited to mm/s, leading to measurement times of many hours for large surfaces. Due to the orthogonal setup, measurement uncertainty rapidly increases with surface inclination (Figure 1.8, right).

Experimental methods

Swing-arm profilometry (Anderson and Burge, 1995; Callender et al., 2006) is able to scan a spherical surface using only two axes of rotation whilst also keeping the probe perpendicular to the best-fit-sphere. Slope measurement (Qian et al., 1995; Van der Beek et al., 2002), slope difference measurement (Geckeler and Weingärtner, 2002) and curvature scanning (Schulz and Weingärtner, 2002; Machkour et al., 2006) determine surface form from integrating surface slope and local curvature, respectively. The latter two properties can be measured independently from an outside reference. Fringe projection calculates the local surface slope and integrates that to height data from a fringe pattern that is deformed after reflecting on a surface under test (Knauer et al., 2004). Advantages and disadvantages of these methods are further discussed in Appendix B and throughout the conceptual design in Chapter 2.

Conclusion

No single method yet incorporates the previously mentioned five characteristics desired for measuring single piece high precision freeform surfaces as described in section 1.1.3. This lack of a suitable metrology method is currently holding back the common application of high-precision aspherical and freeform optics. 

1.2 The NANOMEFOS project

1.2.1 Objective

The objective of the NANOMEFOS[1] project is to realize a new measurement machine prototype for measuring high-end single piece freeform surfaces as defined in section 1.1.3. Hereto, the design goals for this method are formulated as follows.

  1. Universal measurement of:

a)      Flat, spherical, aspherical, freeform and off-axis surfaces

b)     Convex and concave surfaces 

c)      Transmission and reflection optics

  1. Product dimensions up to ∅500 x 100 mm
  2. Measurement uncertainty 30 nm (2σ)
  3. Non-contact
  4. Measure form
  5. Fast (~15 min.)

 

1.2 The NANOMEFOS project

 

 

 

Recapitulating from section 1.1.3, the aspherical departure from spherical is not limited, and the PV freeform departure is 5 mm. The global slope ranges from -45° to +90°, and the unidirectional local slope may amount up to 5°. The reflectivity may range from 4% to 100% and surfaces are not (multi-layer) coated. The chosen product dimensions match the capabilities of the TNO facilities (Appendix A). The measurement uncertainty corresponds with diffraction limited performance for the visible spectrum (λ/20). Measuring form requires a point spacing in the order of 1 mm. Aim is to measure a ∅500 mm surface within 15 minutes. The instrument will be placed in a conditioned metrology laboratory, with an expected temperature stability of 20±0.2°C.

 

A few further objectives are optional. During production, the roughness decreases from opaque to optical quality. Measuring rough surfaces, probably with less accuracy, would therefore be a useful option. Measuring small areas with m point spacing would provide valuable information on roughness and waviness. This would result in a single measurement method that can be applied throughout the entire production process, from roughly ground to finely polished, and from μm to m spatial frequency range.

Ideally, the method would be applicable as an on-machine method. This has been considered, but the boundary conditions of existing manufacturing machines are not quite optimal for metrology (e.g. stage constructions, abrasive slurries, temperature and humidity).

            1.2.2   Methods

The focus in this thesis is on the combination of the previously mentioned five properties into one instrument. The measurement volume of ∅500 x 100 mm, combined with an uncertainty of 30 nm, requires a dynamic range that is one to two orders of magnitude larger compared to state-of-the-art 3D CMMs. The non-contact and fast measurement requires a long range optical probe, which does not yet exist. The application of an optical probe in combination with the large surface slopes requires a departure from the orthogonal setup for the metrology as well as the structural loop design.

 

Numerous papers on precision instrument design principles have been published over the years, of which an overview is published in (Schellekens et al., 1998).

Summarizing, the following principles are applied in the design of the new instrument.

Design for repeatability

As in many precision instruments, emphasis is on design for repeatability in combination with calibration. To achieve the desired measurement uncertainty, the surface measurement repeatability should be in the order of a few nanometers rms. 

In this case, repeatability is not as important in positioning as it is in measuring. Since the intended method is non-contact, a position error is not a problem as long as the actual position is measured correctly. This in contrast to for instance a diamond turning machine, where a position error of the tool is immediately copied into the work piece surface.

By applying statically determined design, hysteresis and backlash are minimized. Distortions resulting from varying loads on the structural loop should also be minimized, since these may give rise to hysteresis and other possible measurement errors. This is done by applying separate preload, position and metrology frames. 

Minimize metrology loop error sensitivity

According to the Abbe principle (Abbe, 1890), the measurement systems are to be aligned with the point of interest to avoid sensitivity to angular displacements. Measurement is further to be done as directly to the point of interest as possible, to obtain a short metrology loop and thus exclude possible error sources from the measurement. Thermal sensitivity of the metrology loop is first of all reduced by minimizing internal and external heat sources. Second, materials with low thermal sensitivity, high thermal diffusivity or low thermal expansion are applied to minimize deformation due to thermal influences.

Minimize system dynamics effects

Applying the minimal number of moving axes generally improves machine accuracy. Light and stiff structural loop design minimizes the required actuation force, increases bandwidth and allows for higher speed and shorter measurement time. By aligning the actuators with the centre of gravity of the stages, parasitic angular displacements are minimized.

Traceability

The required uncertainty level can only be obtained by repeatability in combination with calibration. Calibration can be done by either measuring a reference artefact (which is difficult to obtain to the required accuracy in this case), or by characterizing all error sources of the metrology loop, either individually or in groups. Care must thus be taken that the complete metrology loop can be calibrated. By taking the necessary calibration procedures into account early in the design phase, traceability is assured. 

1.3 Thesis outline

 

 

1.3 Thesis outline

The conceptual design is explained in Chapter 2. Here the cylindrical machine concept is explained after some conceptual considerations and evaluations of existing principles. Next the error budget is determined and a machine design overview is given. The air bearing motion system, with the vertical stage directly aligned to a vertical base plane, is explained in Chapter 3. The overall concept is explained first, after which the design, realization and testing of each motion axis is shown. The chapter is concluded with test results of the assembled motion system. In Chapter 4 the metrology system is explained, starting with the metrology concept, followed by the design, analysis and realization of the interferometry system. Next, the design, realization and testing of the Silicon Carbide metrology frame and multi-probe spindle error measurement are explained. This chapter ends with test results of the metrology system and motion system assembly. Chapter 5 explains the optical probe design. The differential confocal measurement method is explained first, followed by the optical design with the integration of an interferometer and PSD. Next, the optomechanical design is shown, followed by the realization and stand-alone test results. Chapter 6 explains the functionality of the machine control electronics and software. The safety precautions implemented are also addressed. The test results of the machine prototype are shown in Chapter 7, from stability measurements to full surface measurements. Calibration is addressed, and the chapter closes with an achievable uncertainty estimation. Final conclusions and recommendations will be drawn in Chapter 8.



[1] NANOMEFOS is an acronym for Nanometer Accuracy NOn-contact MEasurement of Freeform Optical Surfaces. This project was a continuation of the research done in the authors’ M.Sc. graduation (Henselmans, 2003). It is a collaboration of TNO Science & Industry, Technische Universiteit Eindhoven and the Netherlands

Metrology institute Van Swinden Laboratory. Subsidy was provided by the SenterNovem IOP Precision Technology program of the Dutch Ministry of Economic Affairs. The machine has been realized at the TU/e GTD workshop.

 

Figure 1.3: Freeform optics in the Polaroid SX-70

posted @ 2018-04-09 16:09  lhmchn  阅读(321)  评论(0编辑  收藏  举报