4
28 th ICPIG, July 15-20, 2007, Prague, Czech Republi Time- and spatially-resolved characterization of electrical discharge machining plasma A. Descoeudres 1* , Ch. Hollenstein 1 , G. Wälder 2 , R. Demellayer 2# and R. Perez 2 P 1 Ecole Polytechnique Fédérale de Lausanne (EPFL), Centre de Recherches en Physique des Plasmas, 1015 Lausanne, Switzerland 2 Charmilles Technologies SA, 1217 Meyrin, Switzerland (*present address: CERN, 1211 Geneva 23, Switzerland ; # present address: EIG, 1202 Geneva, Switzerland) Plasma created during electrical discharge machining (EDM) is investigated with fast imaging and with time- and spatially-resolved optical emission spectroscopy. After the breakdown, the plasma develops very fast (< 50ns) and then remains stable. The plasma excites a broad volume around the electrode gap. Typical spectra show a strong H α and continuum radiation, with many lines emitted by impurities coming from electrodes materials. The plasma contamination from the electrodes is mostly concentrated in their vicinity. The electron density reaches 2·10 18 cm -3 at the beginning of the discharge. This extreme density causes merging of lines, strong Stark broadening and shift of the H α line. The density decreases afterwards rapidly with time. The electron temperature remains roughly constant around 0.7eV. The low temperature, the high density measured and other spectroscopic evidences prove that the EDM plasma is non-ideal (Γ 0.45). 1. Introduction Electrical Discharge Machining (EDM) is a well- known machining technique since more than fifty years. Its principle is to use the eroding effect on the electrodes of successive electric spark discharges created in a dielectric liquid. EDM is nowadays widely-used in a large number of industrial areas. Nevertheless, few studies have been done on the discharge itself and on the plasma created during this process, mainly due to the complex physics involved in EDM [1-2]. Further improvements of EDM, especially for micro-machining, require a better control and understanding of the discharge and of its interaction with the electrodes. A better comprehension of the sparking process will also reduce problems related to its stochastic nature. Furthermore, experimental characterization of the EDM plasma is lacking. Some spectroscopic measurements have been made but remain very incomplete [3]. The EDM plasma is experimentally difficult to investigate (small size, weak light intensity, poor reproducibility). These difficulties are the main cause for the lack of experimental data, which are nevertheless essential as inputs and for the validation of numerical models. In this paper, the EDM plasma is systematically investigated with endoscope imaging, and with time- and spatially-resolved optical emission spectroscopy. 2. Experimental setup We use a small and versatile die-sinking EDM machine, equipped with a generator of the Roboform type from Charmilles Technologies. The electrodes are cylindrical, with a diameter of 5mm. In order to better control the localization of the sparks, their tips are conical. The servo-controlled movement of the upper electrode is only vertical. We use pure water or mineral oil as dielectric, copper electrodes and steel workpieces. Plasma imaging is made with a 30000 fibres endoscope, directly immersed in the dielectric. The endoscope is equipped with a small lens at its tip, in order to have a magnification of the plasma region. The image is detected with a Princeton Instruments PI-MAX camera (intensified CCD camera, < 2ns gating, 16 bit, 1024 x 1024 pixels). Fig. 1: Experimental setup for time- and spatially- resolved spectroscopy. Figure 1 shows the experimental setup for time- and spatially-resolved optical emission spectroscopy. In order to have a spatial sampling of the emitted light, the magnified plasma image captured with the endoscope is projected onto an in- line array of 16 fibres. Each fibre collects the light coming from a different zone of the emitting region. The fibres bundle brings the sampled light into a 0.75m imaging spectrograph ARC SP750i, equipped with three gratings (150, 600 and 1800g/mm). The 16 different spectra are recorded simultaneously with the ICCD camera. Due to the small size of the

Time- and spatially-resolved characterization of ...icpig2007.ipp.cas.cz/files/download/cd-cko/ICPIG2007/pdf/T16.pdf · Time- and spatially-resolved characterization of electrical

  • Upload
    others

  • View
    38

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Time- and spatially-resolved characterization of ...icpig2007.ipp.cas.cz/files/download/cd-cko/ICPIG2007/pdf/T16.pdf · Time- and spatially-resolved characterization of electrical

28th ICPIG, July 15-20, 2007, Prague, Czech Republic

Time- and spatially-resolved characterization of electrical discharge machining plasma

A. Descoeudres1*, Ch. Hollenstein1, G. Wälder2, R. Demellayer2# and R. Perez2

P

1 Ecole Polytechnique Fédérale de Lausanne (EPFL), Centre de Recherches en Physique des Plasmas, 1015 Lausanne, Switzerland

2 Charmilles Technologies SA, 1217 Meyrin, Switzerland (*present address: CERN, 1211 Geneva 23, Switzerland ; #present address: EIG, 1202 Geneva, Switzerland)

Plasma created during electrical discharge machining (EDM) is investigated with fast imaging and with time- and spatially-resolved optical emission spectroscopy. After the breakdown, the plasma develops very fast (< 50ns) and then remains stable. The plasma excites a broad volume around the electrode gap. Typical spectra show a strong Hα and continuum radiation, with many lines emitted by impurities coming from electrodes materials. The plasma contamination from the electrodes is mostly concentrated in their vicinity. The electron density reaches 2·1018cm-3 at the beginning of the discharge. This extreme density causes merging of lines, strong Stark broadening and shift of the Hα line. The density decreases afterwards rapidly with time. The electron temperature remains roughly constant around 0.7eV. The low temperature, the high density measured and other spectroscopic evidences prove that the EDM plasma is non-ideal (Γ ≅ 0.45).

1. Introduction

Electrical Discharge Machining (EDM) is a well-known machining technique since more than fifty years. Its principle is to use the eroding effect on the electrodes of successive electric spark discharges created in a dielectric liquid. EDM is nowadays widely-used in a large number of industrial areas. Nevertheless, few studies have been done on the discharge itself and on the plasma created during this process, mainly due to the complex physics involved in EDM [1-2]. Further improvements of EDM, especially for micro-machining, require a better control and understanding of the discharge and of its interaction with the electrodes. A better comprehension of the sparking process will also reduce problems related to its stochastic nature.

Furthermore, experimental characterization of the EDM plasma is lacking. Some spectroscopic measurements have been made but remain very incomplete [3]. The EDM plasma is experimentally difficult to investigate (small size, weak light intensity, poor reproducibility). These difficulties are the main cause for the lack of experimental data, which are nevertheless essential as inputs and for the validation of numerical models.

In this paper, the EDM plasma is systematically investigated with endoscope imaging, and with time- and spatially-resolved optical emission spectroscopy.

2. Experimental setup

We use a small and versatile die-sinking EDM machine, equipped with a generator of the Roboform type from Charmilles Technologies. The electrodes

are cylindrical, with a diameter of 5mm. In order to better control the localization of the sparks, their tips are conical. The servo-controlled movement of the upper electrode is only vertical. We use pure water or mineral oil as dielectric, copper electrodes and steel workpieces.

Plasma imaging is made with a 30000 fibres endoscope, directly immersed in the dielectric. The endoscope is equipped with a small lens at its tip, in order to have a magnification of the plasma region. The image is detected with a Princeton Instruments PI-MAX camera (intensified CCD camera, < 2ns gating, 16 bit, 1024 x 1024 pixels).

Fig. 1: Experimental setup for time- and spatially-resolved spectroscopy.

Figure 1 shows the experimental setup for time-

and spatially-resolved optical emission spectroscopy. In order to have a spatial sampling of the emitted light, the magnified plasma image captured with the endoscope is projected onto an in-line array of 16 fibres. Each fibre collects the light coming from a different zone of the emitting region. The fibres bundle brings the sampled light into a 0.75m imaging spectrograph ARC SP750i, equipped with three gratings (150, 600 and 1800g/mm). The 16 different spectra are recorded simultaneously with the ICCD camera. Due to the small size of the

Page 2: Time- and spatially-resolved characterization of ...icpig2007.ipp.cas.cz/files/download/cd-cko/ICPIG2007/pdf/T16.pdf · Time- and spatially-resolved characterization of electrical

28th ICPIG, July 15-20, 2007, Prague, Czech Republic

light emitting region, this arrangement is the easiest way to perform the spatial sampling.

More details about the EDM machine and the diagnostics can be found in [4-6].

3. Plasma imaging 3.1. Beginning of the discharge

Figure 2 shows fast imaging of the first 300ns of the discharge, obtained with the ICCD camera. This time interval corresponds to the emission of a bright initial light related to the first fast current rise [4]. This interval is thus important in the plasma evolution: it is during the very beginning of the discharge that the plasma forms and that a high amount of energy is brought into it.

Fig. 2: Fast imaging of the discharge beginning, along with current and voltage evolution (50ns exposure, variable delay; 6A, water). The images are normalized in intensity. Each image is obtained during a different discharge.

Due to gating electronics and cables delay, the

first image can only be acquired 50ns after the breakdown. Between 100 and 200ns after the breakdown, the current reaches its maximum. The light intensity is also at its maximum. However, on the plasma images, no clear evolution is visible from 50 to 300ns: size and geometry remain constant. Images obtained with delays longer than 300ns are not shown here, but they are similar to those of the figure 2. This shows that the plasma develops very rapidly (within 50ns) after the breakdown, and then remains quite stable during the whole discharge.

The light emitting region is generally round or oval. Its diameter increases with the discharge current, from 50 to 400μm for a current increase

from 6 to 48A. But even for the same discharge parameters, plasma images are poorly reproducible. The plasma, and so the shape and size of the emitting region recorded on the images, depend strongly on the electrode surface state, which constantly evolves during machining. However, the light mostly originates from a broader region than the gap itself, i.e. the discharge excites a broad volume around the electrode gap (gap ≅ 10-100μm).

3.2. Post-discharge

We have also investigated the afterglow of the discharge. The current and the voltage are shut down and drop rapidly to zero. Simultaneously, the total light intensity also drops fast. But there is still a weak slowly-decaying light emission, typically until 400μs after the end of the discharge. Figure 3(a) shows an image acquired directly after a discharge, and figure 3(b) shows the optical spectrum of the post-discharge light.

Fig. 3: Incandescence of the removed particles after a discharge. (a) Image of a single post-discharge (100μs exposure; 12A, 50μs, oil). (b) Optical emission spectrum, with a 2300K blackbody fit. The light measurement is time integrated over thousands of post-discharges (50μs exposure, 150g/mm grating; 16A, 50μs, oil).

As shown in figure 3, the weak light emitted after

the discharge is due to particles of heated metal. These particles come from the molten metal pool created in the workpiece during the discharge, and are then removed from the workpiece and ejected in the dielectric when the discharge is shut down. We see their path and not only luminous dots, because they move during the camera exposure. Spectroscopy shows that this post-discharge light emission is close to a blackbody radiation, which confirms that the emitters are heated metal particles. The spectrum in figure 3(b) is quite noisy, due to the low intensity of the afterglow emission. Fitting the afterglow spectra with Planck's law, the temperature of the emitters is found to be around 2200K (± 100K). Since the melting point of steel is ~ 1700K, the particles are still in a liquid state in the very beginning of the post-discharge.

Page 3: Time- and spatially-resolved characterization of ...icpig2007.ipp.cas.cz/files/download/cd-cko/ICPIG2007/pdf/T16.pdf · Time- and spatially-resolved characterization of electrical

28th ICPIG, July 15-20, 2007, Prague, Czech Republic

The particle size can be measured on the images. We found that the largest particles have a diameter of ~ 30μm, which is consistent with previous work on spark-eroded particles [7]. The particle speed can also be estimated from images. The maximum speed is around 3m/s.

4. Plasma optical emission spectroscopy 4.1. Typical spectrum

The dominant line is the Balmer Hα line emitted by atomic hydrogen, which comes from the cracking of the dielectric molecules. Some lines of atomic carbon and C2 molecules (Swan system) are also visible. This indicates that the organic molecules of oil are almost completely cracked by the discharge.

The plasma is contaminated by impurities: several lines of atomic copper from the electrode are present, along with many lines originating from atomic iron, chromium and carbon of the removed material of the steel workpiece.

An intense broadband continuum radiation is also observed. It is probable that this continuum is due to free-bound transitions (recombination processes). Free-free radiation is also another plausible source. Since molecules generally emit a broadband spectrum, molecules or fragments of molecules from the dielectric certainly also participate to the continuum [8].

Fig. 4: Typical spectrum (12A, 2μs, oil, 150g/mm grating)

4.2. Plasma contamination

The spectral region around 520nm is interesting for a qualitative characterization of the plasma contamination, because of the presence of three chromium lines at 520.45, 520.6 and 520.84nm, and of one copper line at 521.82nm. With a copper electrode and a stainless steel workpiece, the Cu line is emitted by particles coming from the electrode and the Cr lines are emitted by particles coming from the workpiece. Figure 5 shows spatially-resolved spectra (but time-integrated) along the vertical axis of this spectral region.

An asymmetry of the plasma contamination is clearly visible. Near the Cu electrode, the Cu line is more intense than the Cr lines. On the other hand, the Cr lines are more intense than the Cu line near the steel workpiece. Thus, even if Cu and Cr lines are present everywhere in the plasma, we see that each electrode contaminates the plasma mostly in a region close to itself. Unsurprisingly, no asymmetry is visible along the horizontal axis (measurement not shown here).

Fig. 5: Vertical asymmetry of the plasma contamination (6A, 100μs, water, 150g/mm grating, spectra normalized to Cu line at 521.8nm).

4.3. Electron density and electron temperature

To calculate the electron density ne, we use full width at half-maximum (FWHM) and shift measurements of the hydrogen Hα line [4]. The intensities of three copper lines (510.5, 515.3 and 521.8nm) can be used for electron temperature Te calculation with the Boltzmann plot method (assuming LTE) [4].

As shown in figure 6, the Hα line is extremely Stark-broadened and shifted at the beginning of the discharge, sign of a very high electron density. The evolution of ne profiles, deduced from these spectra, is shown in figure 7.

Fig. 6: Time- and spatially-resolved spectra of Hα (12A, 50μs, water; time resolution 2μs, 600g/mm grating): (a) 0 to 2μs; (b) 4 to 6μs; (c) 36 to 38μs.

The density is extremely high, especially at the

beginning of the discharge (~ 2·1018cm-3 during the first microsecond). Then it decreases with time, remaining nevertheless above 1016cm-3 after 50μs.

Page 4: Time- and spatially-resolved characterization of ...icpig2007.ipp.cas.cz/files/download/cd-cko/ICPIG2007/pdf/T16.pdf · Time- and spatially-resolved characterization of electrical

28th ICPIG, July 15-20, 2007, Prague, Czech Republic

During the whole discharge, the density is slightly higher in the plasma center.

Fig. 7: Evolution of the electron density profiles (12A, 50μs, water, time resolution 2μs).

The EDM plasma has such a high density

because it is formed from a liquid (dense medium). In the very beginning of the discharge, the plasma has to overcome the extreme pressure imposed by the dielectric. Then the plasma expands, which results in a decrease in its density [8]. But during the whole discharge, the density remains high due to the constant pressure imposed by the surrounding liquid.

The electron temperature found from time- and spatially-resolved spectra of Cu lines is around 0.7 ± 0.15eV (~ 8100 ± 1750K). Te is slightly higher at the beginning of the discharge, but rather constant within the margin of error during the whole discharge. Profiles show also that Te is homogeneous in the whole plasma. The temperature value of 0.7eV is a low electron temperature, but consistent with previous studies on EDM plasmas [2, 3] and other similar plasmas [8-10]. The plasma is very dense, and thus electrons rapidly lose energy by numerous collisions. This leads to a low electron temperature.

5. Non-ideality of the plasma

Knowing the typical density and temperature of the EDM plasma, its coupling parameter Γ can be calculated [4]. This parameter represents the ratio of the Coulomb interaction divided by the thermal interaction in the plasma. We found that Γ is around 0.33 (up to 0.45 during the first microsecond). Thus, EDM discharges produce cold and dense plasmas which are weakly non-ideal. In these plasmas, the Coulomb interactions between the charged particles are of the same order as the mean thermal energy of the particles, which produces coupling phenomena.

In addition to calculation of Γ, spectroscopic results confirm that the plasma is non-ideal. The strong broadening and shift of the Hα

line and its

asymmetric shape and complex structure [4], the

absence of the Hβ line [4], and the merging of spectral lines (as shown in figure 8), are indeed typical of non-ideal plasmas [11].

Fig. 8: Merging of the Fe, Cu and Cr atomic lines during the first microsecond (12A, water; time resolution 200ns, 600g/mm grating, spectra normalized to the Cu line at 521.8nm).

6. Conclusion

The experimental investigation of the EDM plasma is difficult. Nevertheless, this work gives a first insight into fundamental aspects of this plasma. The EDM plasma has extreme physical properties (especially its ne), and the physics involved is astonishingly complex (non-ideal effects).

7. Acknowledgment

This work is funded by the Swiss Federal Research grant TopNano 21, project no 5768.2.

8. References

[1] F. Van Dijck, PhD Thesis, Catholic University of Leuven, Belgium (1973).

[2] P.T. Eubank et al., J. Appl. Phys., 73 (1993) 7900.

[3] K. Albinski et al., Plasma Sources Sci. Technol., 5 (1996) 736.

[4] A. Descoeudres et al., J. Phys. D: Appl. Phys., 37 (2004) 875.

[5] A. Descoeudres et al., J. Phys. D: Appl. Phys., 38 (2005) 4066.

[6] A. Descoeudres, PhD Thesis n°3542, Ecole Polytechnique Fédérale de Lausanne, Switzerland (2006).

[7] J. Carrey et al., J. Appl. Phys., 95 (2004) 823. [8] P. Barmann et al., J. Phys. D: Appl. Phys., 29

(1996) 1188. [9] P. Frayssines et al., J. Phys. D: Appl. Phys.,

35 (2002) 369. [10] A. Escarguel et al., J. Quant. Spectrosc.

Radiat. Transfer, 64 (2000) 353. [11] G. Kobzev et al., Transport and optical

properties of nonideal plasma, Plenum Press, New York (1995).