Sie sind auf Seite 1von 7

Applied Surface Science 208209 (2003) 285291

Physical chemistry of the femtosecond and nanosecond lasermaterial interaction with SiC and a SiCTiCTiB2 composite ceramic compound
Pascale Rudolph*, Klaus-Werner Brzezinka, Rolf Wasche, Wolfgang Kautek
Federal Institute for Materials Research and Testing Unter den Eichen 87, D-12205 Berlin, Germany

Abstract The interaction of nanosecond laser pulses in the ultraviolet wavelength range and femtosecond laser pulses in the near-infrared region with the semiconductor SiC and the composite compound SiCTiCTiB2 was investigated. Surface analytical techniques, such as XPS, depth prole (DP), and micro-Raman spectroscopy (m-RS) were used to identify the chemical changes between untreated and laser-treated areas. Single-pulse irradiation led to material modications in the condensed state in most instances. Multi-pulse results differed depending on the pulse duration. Crystal structure changes were observed as a consequence of laserinduced melting and resolidication. In air contact all components underwent oxidation reactions according to thermodynamic expectations. Exceptions were observed under exclusion of oxygen, SiC was reduced to elemental Si. # 2002 Elsevier Science B.V. All rights reserved. PACS: 78.47; 06.60.J; 81.05.J; 78.30; 79.60
Keywords: Femtosecond techniques; Nanosecond techniques; Ceramics; Raman spectra/condensed matter; XPS/surface analysis

1. Introduction Silicon carbide plays an important role in many industrial applications because of its hardness, high melting temperature, chemical and thermal resistance, combined with its low weight [1]. The composite ceramic compound contains silicon carbide (SiC), titanium carbide (TiC) and titanium diboride (TiB2) and was designed for tribological issues. Addition of TiC and TiB2 to the SiC matrix reduces the wear rate [2]. Mechanical machining of these materials is
Corresponding author. Tel.: 49-30-8104-3845; fax: 49-30-8104-1827. E-mail address: pascale.rudolph@bam.de (P. Rudolph).
*

difcult due to their hardness up to 22 GPa [3]. A solution may be the application of short and ultra-short laser pulses to avoid mechanical tool wear and minimize thermal and mechanical stress. Laser ablation is associated with physicochemical changes. In this study, some of them were investigated by surface analytical techniques, such as changes in crystal structure after melting, and resolidication, and oxidation. Morphologies and threshold uences were determined. 2. Experimental Silicon carbide (SiC) and the composite ceramic compound (SiCTiCTiB2) are sintered ceramics.

0169-4332/02/$ see front matter # 2002 Elsevier Science B.V. All rights reserved. doi:10.1016/S0169-4332(02)01356-9

286

P. Rudolph et al. / Applied Surface Science 208209 (2003) 285291

They were produced at 2180 8C from a powder mixture of the compounds (H.C. Starck) mixed with sinter additives as Alnovol (Hoechst, PN846) and Al-TriIsopropylat (Merck-Schuchardt) [4]. The silicon carbide powder (Grade UF-15 ``Premix'') represents mainly the a-polytype. The composite contains silicon carbide (SiC), titanium carbide (TiC) and titanium diboride (TiB2) in the relation 44:42:14 mol%. The untreated SiC surface shows grain boundaries and graphitic additives in the order of $5 mm. In the SiCTiCTiB2 composite, SiC and TiB2 are present as cystallites in the ceramic matrix, whereas TiC stays as cement in between. The grain size is approximately 510 mm. Before laser treatment, ultrasonic cleaning in methanol was done. A frequency quadrupled Nd:YAG laser system (Spectron Laser Systems, SL850) was used at the wavelength of l 266 nm for the nsablation experiments. The output energy per pulse was determined by a pyroelectric detector (BESTEC, PEM 521). The pulse duration (t 10 ns) was measured with a fast photodiode and an oscilloscope. The repetition rate was 2.5 Hz. The pulses were focused on the sample with a lens of 50 mm focal length yielding a beam spot of f % 42 mm. In the sub-picosecond range, a fs-Ti:sapphire laser system (Spectra Physics, Spitre) emitting pulses of t 130 fs at l 800 nm was used. Pulse energies were measured by a pyroelectric detector (BESTEC, PM 200). Pulse durations were checked with a scanning autocorrelator (APE Berlin, PulseScope). The repetition rate could be varied from 1to 1000 Hz. The pulses were focused on the sample with a 60 mm focal length lens to a beam spot of f % 34 mm. In both experimental setups, the direction of the incident laser beam was perpendicular to the treated surface and the sample was mounted on a motorized xyz translation stage. An optical microscope was used to investigate the cavity diameters and the hole depths. A scanning electron microscope (SEM) equipped with a cold-eld electron emission cathode (Hitachi, S-4100, accelerating voltage 10 kV) served to study the surface morphology. XPS (Surface Science Instruments, SSX-100-SProbe) with a spot of (250 mm 1000 mm) was performed with a monochromatic aluminium Ka X-ray source. To determine the binding conditions, the analysis was made with an energy resolution for selected

elements of DEkin $ 0:1 eV. Depth proles (DPs) were performed by Ar ion sputtering. Micro-Raman spectroscopy (m-RS) gives information about the crystal structure. The spectrometer (Dilor-XY) with Ar ion-laser excitation detected in the spectral range of 1001800 cm1 and contained a nitrogen cooled CCD camera as multi canal detector in 1808 backscattering geometry. The electrical eld vector of the linear polarized laser beam was in plane with the sample surface. The laser radiation with an output power of 1025 mW was focused with a BH2Olympus Microscope on the sample (focus diameter $2 mm, intensity <2.5 mW/mm2). 3. Results and discussion The semiconductor silicon carbide (SiC) has a direct band gap of Eg 3:0 eV. Therefore, a linear absorption of the ultraviolet nanosecond laser with the wavelength of l 266 nm and a photon energy of EPhoton(ns) 4.7 eV is sufcient to cross the band gap. In the case of the infrared femtosecond laser (l 800 nm), two photons of the energy EPhoton(fs) 1.6 eV are needed. Equivalent data are not available for the composite ceramic compound SiCTiCTiB2. The application of one laser pulse in both duration regimes did not result in ablation of all specimens, only in a modication. It needed a nite number of pulses to detect a macroscopic removal of material. This so-called incubation is often observed in laser micromachining. Around the ablated areas also modication was detectable. To determine the ablation or modication threshold uences, diameters were evaluated with an optical microscope. For N 100 pulses, the squared diameters D2 as a function of ln(F) yielded a linear relationship [57]. An example for this technique is shown for the nanosecond laser application on SiC in Fig. 1. The extrapolation to D2 0 delivered the ablation and modication threshold uence Fth and Fmod, respectively: Fth(SiC, ns) 0.63 J/cm2, Fmod(SiC, ns) 0.25 J/cm2. An equivalent procedure gave the data for the fs-ablation: Fth(SiC, fs) 1.79 J/cm2, Fmod(SiC, fs) 0.43 J/cm2. A determination of the modication threshold uence of the composite ceramic compound was not possible because the laser beam diameters were almost of the same dimension as the inhomogeneities. Ablation thresholds are

P. Rudolph et al. / Applied Surface Science 208209 (2003) 285291

287

Table 1 Survey of the threshold uences Fx (J/cm2) of silicon carbide (SiC) and the composite ceramic compound (SiCTiCTiB2) with nanosecond and femtosecond laser systems SiC ns Fmod Fth 0.25 0.63 fs 0.43 1.79 SiCTiCTiB2 ns 1.38 fs 1.68

Fmod modication threshold uence for N 100, Fth ablation threshold uence for N 100 in air.

Fig. 1. Nanosecond laser ablation of silicon carbide in air. l 266 nm, t 10 ns, N 100, f 50 mm, linearly polarized. Squared diameters D2 of the modied and ablated areas in dependence on the applied laser uences ln(F).

Fth(SiCTiCTiB2, ns) 1.38 J/cm2 and Fth(SiC TiCTiB2, fs) 1.68 J/cm2, respectively (see also Table 1). A comparison of the morphological changes after the application of N 100 nanosecond pulses

Fig. 2. Comparison of nanosecond and femtosecond laser ablation on silicon carbide and the composite ceramic compound SiCTiCTiB2 in air. ns: l 266 nm, t 10 ns, N 100, F0 3x Fth. (a) F 1:9 J/cm2, (c) F 4:2 J/cm2. fs: l 800 nm, t 130 fs, N 100, F0 3x Fth. (b) F 5:6 J/cm2, (d) F 7:0 J/cm2.

288

P. Rudolph et al. / Applied Surface Science 208209 (2003) 285291

Table 2 Oxidation reactions of SiC, TiC, and TiB2 [9] Oxidation reactions SiC 2O2 3 SiO2 CO2 SiC 3/2O2 3 SiO2 CO TiC 2O2 3 TiO2 (Rutil) CO2 TiC 3/2O2 3 TiO2 (Rutil) CO 2TiC 5/2O2 3 Ti2O3 2CO 3TiC 4O2 3 Ti3O5 4TiC 11/2O2 3 Ti4O7 4CO TiB2 5/2O2 3 TiO2 (Rutil) B2O3 2TiB2 9/2O2 3 Ti2O3 2B2O3 3TiB2 7O2 3 Ti3O5 3B2O3 4TiB2 19/2O2 3 Ti4O7 4B2O3
a

DGf (kJ)a 298 K 589 614 551 564 539 547 553 719 722 720 720 1200 K 513 565 476 515 514 517 518 579 586 585 584

DGf connected with an oxygen amount of 1 mol.

for both ceramic materials is depicted in Fig. 2a and c. Fluences were three times the ablation threshold uence of the specic material. The edges are covered with smooth resolidied melt phases. In both cases, substantial debris has been deposited next to the cavity. It appeared brighter due to charging and low conductivity. Micromachining results with N 100 femtosecond pulses are shown in Fig. 2b and d. No melt phases, but oriented periodic structuresso-called ripples occurred. They have a periodicity near the applied laser wavelength of l 800 nm or even smaller and they are oriented perpendicular to the vector of the incident electrical eld. This phenomenon is quite often observed in micromachining with ultra-short pulse durations [8]. All laser treatments in air involved oxidative conversion. Table 2 lists possible oxidation reactions of SiC, TiC, and TiB2 together with free energies of formation at room temperature and at 1200 K [9]. Especially for the titanium oxides, several variations are possible. A quite common species represents the Magneli phase TinO2n1, which was also found via transmission electron microscopy (TEM) at tribologically treated samples [10]. Ti2O3 has a black appearance and might be therefore the dark rings around the laser cavities. Deviations from these thermodynamically expected reactions have been investigated by photoelectron spectroscopy (XPS), depth prole, and micro-Raman spectroscopy.

Fig. 3. Micro-Raman spectra of silicon carbide (SiC) after nanosecond laser ablation. l 266 nm, t 10 ns, N 5, F 4:2 J/cm2. (a) Untreated area, (b) at the edge, and (c) in the center of the cavity.

The m-RS was performed at multi-pulse cavities. Results of a ns-generated cavity on SiC are depicted in Fig. 3. Spectrum (a) derives from an untreated area, (b) from the crater edge, and (c) from its center. In all cases, the determination of the substrate material was possible because m-RS is sensitive up to a depth of several micrometers. The sharp peaks in the region of 750800 cm1 represent the transversal optical modes and those around 9001000 cm1 the longitudinal optical modes of the polycrystalline SiC [11]. Also the bands of the graphitic additives were detectable as the defect or D-peak at 1360 cm1 and the graphite or G-peak at 1585 cm1 [12] especially at the edge (b). Further bands in spectrum (c), one sharp at $520 cm1 and a weak broad around 480 cm1, represent crystalline and amorphous silicon, respectively [13,14]. Obviously, part of the SiC was not oxidized to SiOx and COx (Table 2), but the Si component underwent reduction to the elemental state. The resolidied melt remained nanocrystalline and amorphous. Additionally, the stray light increased in a wide surface area, particularly in the region <500 cm1, which represents the lattice vibrations.

P. Rudolph et al. / Applied Surface Science 208209 (2003) 285291

289

Fig. 4. Micro-Raman spectra of the composite ceramic compound (SiCTiCTiB2) after femtosecond laser ablation (a) and (b) in the center of the cavity with a distance of 2 mm (l 800 nm, t 130 fs, N 5, F 9:8 J/cm2).

Two spectra after fs-application on SiCTiCTiB2 were taken in the center of the cavity (Fig. 4). On a Tirich grain, all resulting bands correspond to rutile (TiO2) [15] (Fig. 5a). The slight shift of the bands

Fig. 6. XPS analysis of the composite ceramic compound (SiC TiCTiB2). Comparison of the Si 2p line for the untreated sample (a) with the surfaces processed with ns (b) and fs laser pulses (c).

in comparison to the rutile single crystal demonstrates the oxidation to non-stochiometric titanium dioxide [16]. The spectrum (b) taken on a Si-rich grain shows unoriented graphite (D/G-peak) and the SiC-bands around 800 and 1000 cm1. There was no signicant melting with fs-pulses, only the partial formation of crystalline surface layers. Further proof of the non-oxidative side reactions gave XPS. This technique has an information depth of
Table 3 Comparison of the XPS results for nano- and femtosecond laser pulses on silicon carbide and the composite ceramic compound (relative atomic percentage for the Si 2p signal) SiC Untreated Si SiC SiO SiO2 68.1 31.9 ns 35.9 64.1 fs 46.9 39.9 13.2 SiCTiCTiB2 Untreated 87.3 12.7 ns 55.5 44.5 fs 43.6 46.8 9.6

Fig. 5. XPS analysis of silicon carbide (SiC). Comparison of the Si 2p line for the untreated sample (a) with the surfaces processed with ns (b) and fs laser pulses (c).

290

P. Rudolph et al. / Applied Surface Science 208209 (2003) 285291

Fig. 7. Depth prole after femtosecond laser ablation of both ceramics (a) on silicon carbide (SiC) and (b) on the composite ceramic compound (SiCTiCTiB2).

only few nanometers. The analysis was performed with a spot size (250 mm 1000 mm) in the middle of a laser-treated area of $6 mm 6 mm to investigate over several overlapping laser cavities. A comparison of the Si 2p line for the untreated sample with the ns- and fs-laser processed surfaces is given for SiC in Fig. 5 and for SiCTiCTiB2 in Fig. 6, respectively. The accompanying relative atomic percentages for the Si 2p line are listed in Table 3. In both cases, part of the untreated surface contained silicon oxide (Fig. 5a and 6a). The major compound represented SiC with a binding energy of <101 eV [17]. After ns-treatment of SiC, almost exclusively oxide, represented by a band at 102 eV (silicates) and at 103 eV (SiO2) [18] was detected (Fig. 5b). Another interpretation for these binding energies could be the formation of silicon oxycarbides, SixCyOz [19]. In contrast, the fs-processing left the substrate partially unoxidized (Fig. 5c). For SiCTiCTiB2, the XPS analysis gave different results (Fig. 6): laser treatment in both duration regimes yielded about 50% elemental Si (100 eV; Fig. 6b and c). The surface interestingly had not been oxidized by the ns-laser, only SiC has been reduced to Si. With ns-pulses, a DP up to 200 nm yielded no changes in the relative atomic percentages for both materials. One has to conclude that the conversion layer is much thicker and of the order of the heataffected zone ($mm). However, the fs-laser treatment

resulted in a much thinner conversion lm of <130 nm for SiC and <240 nm for SiCTiCTiB2 (Fig. 7). The crossing point of the oxide and non-oxide parts can dene the interface between reaction zone and bulk. This suggests that the heat-affected zone is at least as extended as the chemically conversed region. 4. Conclusion The ablation thresholds of the ns-UV pulse treatment of the semiconductor SiC was Fth 0:63 J/cm2 and that of the composite ceramic compound SiC TiCTiB2 Fth 1:38 J/cm2. In contrast to this, the near-IR fs-processing yielded higher values of $1.7 J/ cm2 independent of the chemical composition. This behaviour is typical for sub-picosecond interactions with transparent materials, where the optical properties become relatively irrelevant due to multi-photon processes. Melted regions were converted to resolidied nanocrystalline and amorphous phases. All components are expected to convert to their oxides when laser-processed in air contact according to thermodynamics. This holds for Ti which forms e.g. rutile (TiO2) or black non-stoichiometric titanium oxide (Magneli phase). SiC showed deviations from the thermodynamic behaviour. Obviously, kinetic reasons excluded the reactant oxygen so that carbon components either in the carbide or the graphitic sinter

P. Rudolph et al. / Applied Surface Science 208209 (2003) 285291

291

additives could lead to a reduced Si species. The reduction of SiC and/or the oxidation of carbon were relatively faster than the formation of a SiOx phase. Acknowledgements Partial nancial support was provided by the German Ministry for Research and Technology (BMBF) in the framework of LASER 2000 (Laserinduzierte Fertigungsverfahren, Verbund projekt ABLATE, #13N 7048/7). The authors would like to thank Mrs. B. Straub, Mrs. S. Benemann, and Mr. D. Treu (BAM Laboratory VIII.23 ``Surface and Thin Film Analysis'') for taking SEM pictures and XPS analysis, and also Mr. R. Yarim (BAM Laboratory V.41 ``Structural Ceramic and Ceramic Components'') for technical assistance. References
[1] A.W. Weimer (Ed.), Carbide, Nitride and Boride Materials Synthesis and Processing, Chapman & Hall, London, 1997. [2] R. Wasche, D. Klaffke, Tribol. Lett. 5 (1998) 173. [3] R. Wasche, T. Rabe, Ceram. Trans. 99 (1998) 537.

[4] J.P. Singh, N.P. Bansal, E. Ustundag (Ed.), Ceramic Matrix Composites VI, Ceramic Transactions, The American Ceramic Society, vol. 124, 2001, p. 185. [5] J.M. Liu, Opt. Lett. 7 (1982) 196. [6] J. Jandeleit, A. Horn, R. Weichenhain, E.W. Kreutz, R. Poprawe, Appl. Surf. Sci. 127129 (1998) 885. [7] J. Bonse, P. Rudolph, J. Kruger, S. Baudach, W. Kautek, Appl. Surf. Sci. 154155 (2000) 659. [8] J. Bonse, Ph.D. Thesis, TU Berlin, 2001. [9] JANAF Tables, J. Phys. Chem. Ref. Data 14, Suppl. 1 (1985) (1988 update). [10] M. Woydt, A. Skopp, I. Dorfel, K. Witke, Wear 218 (1998) 84. [11] K. Gohlert, G. Irmer, L. Michalowsky, J. Monecke, Mat.Wiss. u. Werkstofftech. 21 (1990) 108. [12] C. Palma, M.C. Rossi, C. Sapia, E. Bemporad, Appl. Surf. Sci. 138139 (1999) 24. [13] S. Veprek, F.-A. Sarott, Z. Iqbal, Phys. Rev. B 36 (1987) 3344. [14] W.P. Acker, B. Yip, D.H. Leach, R.K. Chang, J. Appl. Phys. 64 (1988) 2263. [15] J.C. Parker, R.W. Siegel, J. Mater. Res. 5 (1990) 1246. [16] G.M. Begun, C.E. Bamberger, Appl. Spectr. 43 (1989) 134. [17] B. Mitu, G. Dinescu, M. Dinescu, A. Ferrari, M. Balucani, G. Lamedica, A.P. Dementjev, K.I. Maslakov, Thin Solid Films 383 (2001) 230. [18] J. Chastain, Handbook of X-Ray Photoelectron Spectroscopy, Perkin-Elmer Corporation, Physical Electronics Division, Eden Prairie, MN, 1992. [19] S. Baunack, S. Oswald, H.K. Tonshoff, F. von Alvensleben, T. Temme, Fresenius J. Anal. Chem. 365 (1999) 173.

Das könnte Ihnen auch gefallen