Sei sulla pagina 1di 10

Full Paper

Interaction Mechanisms between ArO2 Post-discharge and Biphenyl


Walter DalMaz Silva, Thierry Belmonte,* David Duday, Gilles Frache, Cedric Noe l, Patrick Choquet, Henri-Noel Migeon, Ana Maria Maliska

Interactions between a late ArO2 post-discharge and biphenyl (C6H5)2 are studied. Thin lms grown by spin-coating are efciently etched by the post-discharge to get a clean surface after treatment. Thin lms are strongly preferentially oriented. But this orientation does not depend on the way the sample is pretreated. The pretreatment can create cracks and modify the etching rate. The etching of the biphenyl 1 + OH occurs by interaction with the singlet state of H + OH + O (2p P) molecular oxygen O2(a1Dg). [4 2] cycloaddition + O (a ) [4+2] cycloaddition OH is assumed to be the main process leading to ring OH opening. Next, a large variety of compounds + O (a ) OH including alcohols, ketones, acids, and aldehydes O + O O O OH are created. Atomic oxygen does not seem to play a signicant role in the etching process but it CH + CH O + O (a ) functionalizes the biphenyl by creating alcohol OH +H groups. O
43 2 1 g O O 2 1 g O O 2 2 2 1 g

1. Introduction
Reactive oxygen species can be produced by electrical discharges. They can be transported by the gas ow downstream the plasma to a neutral region, the postdischarge, which is well known but still the subject of many

W. DalMaz Silva, T. Belmonte, C. Noe l Institut Jean Lamour, Department of Chemistry and Physics of , CNRS, Parc de Saurupt, CS Solids and Surfaces, Nancy-Universite 14234, F-54042 Nancy Cedex, France Fax: (33) 3 83 53 47 64; E-mail: thierry.belmonte@mines.inpl-nancy.fr W. DalMaz Silva, A. M. Maliska rio de Materiais, Departamento de Engenharia Laborato nica, Universidade Federal de Santa Catarina, 88040900 Meca polis, SC, Brazil Floriano D. Duday, G. Frache, P. Choquet, H.-N. Migeon Centre de Recherche Public Gabriel Lippmann, 41, rue du Brill, 4422 Belvaux, Luxembourg
Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

studies.[16] Experimental works[79] show that in similar processes, the oxygen atom density has a well-dened maximum as a function of the gas pressure. The oxygen atom density is about several 1014 cm3 when the plasma is operated at a few tens of Pa and microwave powers close to 100 W. When the power rises, the maximum is shifted toward higher pressures. These species are known to modify long molecular chains,[1012] have bactericidal effects[1316] but the interaction mechanisms that occur between them and living micro-organisms are poorly understood.[17,18] To improve our knowledge on the possible reaction pathways between active species produced in ArO2 postdischarges and living materials, we chose to use model molecules,[1928] i.e., molecules with a limited number of chemical bonds. We hope that we shall gain a clear understanding of the way reactive oxygen species modify specically each type of bond. We studied successively the hexatriacontane (C36H74)[2024] and the stearic acid (C18H36O2)[2528] as model molecules.

wileyonlinelibrary.com

DOI: 10.1002/ppap.201100119

207

W. DalMaz Silva et al.

Briey, we could show that the etching of these model molecules occur similarly, the acid group being not modied by the post-discharge. H in CH groups is removed by O to create OH and a radical. This radical reacts with O2 in any state, producing a peroxide. This compound reacts with H available from surrounding molecules to create a hydroperoxide. OH is next removed thermally to give an alkoxy radical whose lability leads to the b-scission of the CC bond. The lightest fragments are gaseous, and removed in the gas phase. The role of the diffusional transport of heavier radicals through the liquid phase and the cross-linking reactions they undergo was also elucidated. Thick lms are much longer to etch than thin lms where this latter process is negligible. Knowing how ArO2 post-discharges react with CC, CH, COOH, rings and next with amine functions, will enable us to describe how amino acids, one of the elemental component of living materials, can be transformed by these active media. Gonzalez et al.[29] studied the interaction of remote plasma at atmospheric pressure with polymers. They suggested that O atoms are responsible for the ring opening of various polymers (polyetheretherketone, polyphenylsulfone, polyethersulfone, and polysulfone). In this work, we want to study another molecule, the biphenyl: (C6H5)2. This molecule, made of two juxtaposed rings, was chosen because its melting point is 342 K, i.e., close to melting temperatures of the HTC and the stearic acid (348 and 342.7 K, respectively). The outline of this work is as follows. After a description of the experimental set-up, transformations of biphenyl are described. Next, results of characterization by various diagnostics are given and an interaction model is nally proposed.

2.45 GHz microwaves by means of a surfaguide wave launcher. The forward microwave power is 150 W. Treatments are carried out at 400 Pa for 1 h in an Ar10 vol.-% O2 mixture, owing at 1 000 sccm. The post-discharge enters a Pyrex tube (28 mm inner diameter) 45 cm downstream the plasma gap. The initial temperature of the substrate is controlled by a device positioned around the Pyrex tube that heats the substrate at a given temperature until the experiment starts. It is 313 2 or 333 2 K at the beginning of the experiment, i.e., when oxygen is added to the plasma. Next, the heating device is turned off and the temperature evolves freely. Indeed, surface reactions release heat and the substrate temperature changes. These changes are not negligible in the case of thin lms and occur during less than 10 min (Figure 2). We studied the inuence of pressure by choosing to treat samples at 400 and 600 Pa. Each experiment was made three times and mean values are presented with a dispersion varying typically between 5 and 20%. Two types of samples were used. For determination of etching rates, a load of about 0.8 g of biphenyl (purity: 98%) is melted and poured into a Teon crucible (1.5 2.2 2.4 cm3) where it solidies. Before its introduction in the reactor chamber, the sample is weighted with a high precision balance (0.01 mg). We know from previous studies[3032] that the crystalline orientation can play a role on the etching rate. Four conditions were selected to determine the inuence of the biphenyl crystalline orientation on the etching rate: after slow (approximately 58 min1) or fast (approximately 158 min1) cooling (from melting temperature), samples were used as such or after annealing at 328 5 K for 60 min in an argon atmosphere at 1 000 Pa. We give in Figure 3 only one corresponding X-ray diffraction pattern because it is the same for all samples. For comparison, the diffraction pattern of the powders used to synthesize these samples is also provided. The various phases are identied using grazing-incidence X-ray diffraction at an incident angle of 38 with Co(Ka) radiation (l 0.178897 nm). For surface characterization, undoped silicon substrates [1 1 cm2 with a (100) orientation250 mm in thickness] are used. They are cut in 200 Single Side Polished wafers provided by Siltronix. They are covered by a silica passive lm. First, they are cleaned in an ultrasonic acetone bath. Next, they are rinsed with

2. Experimental Section
The experimental set-up is depicted in Figure 1. The plasma is created in a 5 mm inner diameter cylindrical quartz tube exposed to

Figure 1. Experimental set-up.


Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 2. Evolution of the sample temperature as function of time at two pressures (400 and 600 Pa) and two initial temperatures (313 and 333 K).

208

DOI: 10.1002/ppap.201100119

Interaction Mechanisms between ArO2 Post-discharge and Biphenyl

meter (Bruker Reex IV MALDI-TOF-MS) and a high resolution Fourier Transform Ion Cyclotron Resonance MS (Ionspec/Varian Explorer FTMS).

3. Results and Discussion


3.1. Etching Rates Contrary to the cases of hexatriacontane and stearic acid, the biphenyl is not topped by a liquid phase when it is modied by the post-discharge.[22,26,27] Degradation products are light enough to be gaseous and removed from the solid phase. As a consequence, the crystalline orientation of the sample can play a role on the rate of the etching process. In Figure 3, XRD analyses show that each sample is strongly textured but no signicant change could be observed by this technique according to the different thermal treatments used to modify the sample (Figure 2). Yet, we show in Figure 4 that the thermal treatment affects the etching rate for given conditions. These changes are attributed to the formation of cracks in the solid. These cracks increase the reactive area of the sample. If the cooling step is fast or if the sample is annealed, we observe an increasing number of cracks. The etching rate is not attributed to a texturing effect, even if it is very pronounced in all cases, but to a change in the specic area of the sample. Consequently, we chose to study only samples cooled slowly and without annealing. In Figure 5, we show that biphenyl is continuously etched by the post-discharge as a function of time. The etching rate increases when the pressure decreases and when the initial temperature increases. It remains inferior to 550 nm s1. The energy released in the solid by surface reactions affects the solid temperature only during the rst minutes of the treatment. Next, the temperature of the solid remains

Figure 3. X-ray diffraction patterns of (a) biphenyl powders used to get samples obtained after melting and cooling at 58 per minute (b) before annealing. XRD patterns for annealed samples are strictly similar to pattern (b).

ethanol and dried. A thin layer of biphenyl is deposited onto each substrate by spin-coating. A melted droplet of biphenyl is deposited on the substrate rotating at 9 000 rpm. The sample thickness is typically about 50 mm as observed by SEM. X-ray photoelectron spectroscopy (XPS) measurements are performed with a Thermo VG Microlab350 spectrometer using a non-monochromated Al Ka and Mg Ka dual anode as X-ray source operated at 300 W and a Spherical Sector Analyser. Survey spectra to identify elements on the surface were recorded in steps of 1 eV at a 100 eV pass energy. High resolution spectra of separate photoelectron lines (C 1s, O 1s, Si 2p, and N 1s) were taken by steps of 0.05 eV at a constant pass energy of 20 eV. The normal operation pressure was 5 109 mbar. The photoelectron take-off angle (TOA) was normal to the surface of the samples. The sample surface covered by the analysis is 2 5 mm2. Spectra processing (atomic concentrations, curve tting, etc.) was done after the removal of a Shirley type background with the CasaXPS software. The atomic concentrations were calculated using the Scoeld relative sensitivity factor provided by the CasaXPS library. Samples are introduced in the spectrometer readily after plasma treatment in order to limit contamination from ambient air storage and directly analyzed. Gas chromatography combined with mass spectrometry (GC/MS) analyses are performed as follows with a Varian GC3800 coupled to a Saturn 2000 ion trap mass spectrometer. Each half sample is extracted with methylene chloride CH2Cl2 (0.6 mL). 1 mL of each solution is injected at 250 8C with a split ratio of 50. The column oven temperature raise from 50 to 320 8C in 18 min. The mass range is set between 80 and 650 amu with an electron impact energy of 70 eV. Laser desorption/ionization mass spectrometry (LDI-MS) analyses were carried out by dropping 1 mL of the sample on the stainless steel target. A Nd:YAG laser (MiniLite II, Continuum, Evry, France) is used to provide a 266-nm UV laser beam (quadrupled frequency). The irradiance conditions were adjusted just above the ionization threshold to prevent laserinduced degradation of the sample. Laser-induced ions were subsequently analyzed by either a Time-Of-Flight mass spectroPlasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 4. Mass variation of the sample after various pre-treatments. Conditions are given in the text.

www.plasma-polymers.org

209

W. DalMaz Silva et al.

Figure 5. Weight loss of samples (not annealed) as a function of time at two pressures (400 and 600 Pa) and two initial temperatures (313 and 333 K). Figure 6. Arrhenius plot of the etching reaction of the biphenyl.

constant. The etching rates follow always the same trend, i.e., a rst-order exponential decay, at least within the accuracy of the raw data. If we assume that temperature and the concentration of the oxygen precursor are constant versus time, we deduce that the etching process is a rstorder process with respect to the biphenyl Ox s ! by products (a)

We can use the concentrations of these species in the present condition to determine the temperature evolution of the rate constant. O2(a1Dg) concentrations are extrapolated from values available in reference.[5] We obtain the Arrhenius plot given in Figure 6. Finally, one nds ! 3 62 8 103 Jmol1 1 1 k m mol s / exp RT

(3)

[w] and [Ox s] are the concentrations of the biphenyl and the oxygen reactive species adsorbed on the biphenyl surface. The etching rate is then equal to d kOx s dt (1)

This value is coherent with activation energies found in thermolyses of endoperoxides.[34,35] 3.2. Characterization Results To clarify the etching processes, thin lms of biphenyl were analyzed after treatment by various characterization techniques. In Figure 7, we show typical spectra of O(1s) and C(1s) peaks. We deduce that a grafting of oxygen occurs from the O(1s) spectrum. On the C(1s) spectrum, we notice a decrease of the CC bonds and an increase of the peak component at 286.9 eV, which can be attributed to CO bonds. We also observe a slight increase of ketone groups and carboxylic acids or esters at higher energies. These latter groups are present in the untreated biphenyl and are likely due to contamination or ageing. To clarify the transformations undergone by the biphenyl in late post-discharge, several analysis techniques were used. Results are provided for a treatment of 3 or 10 min. These times correspond to transformation ratios of the biphenyl close to 30 and 80%. In the second case, the temperature reaches its highest value and we may think that the branching ratios of reactions can be affected, leading to new degradation products.

where k is the rate constant of the etching process (a). Thus    t Dm m0 1 exp t where t 1 : kOx s (2)

By tting results given in Figure 5, we determine the product k[Ox s]. We can go a step further if we assume, as justied hereafter, that the singlet state of molecular oxygen O2(a1Dg) is the main reactive species, and the adsorbed species are proportional to the concentration of this species in the gas phase. This assumption means that the ux of O2(a1Dg) impinging on the surface is given by the kinetic theory of gases (i.e., surface reactions are rate limiting).[33]
Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

210

DOI: 10.1002/ppap.201100119

Interaction Mechanisms between ArO2 Post-discharge and Biphenyl

Figure 8. Positive ion LDI-TOF-MS spectrum [(a) from 85 to 210 amu(b) from 270 to 390 amu] of the biphenyl treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa.

Figure 7. O(1s) (top) and C(1s) (bottom) XPS spectra of the biphenyl treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa.

In Figure 8, we show two details (from 85 to 210 amu in Figure 8a and from 270 to 390 amu in Figure 8b) of the LDITOF-MS positive ion mass spectrum of a sample treated for 3 min. Possible molecules associated with detected ions are listed in Table 1. Peaks are given with a precision of 0.02 amu. We notice three levels of oxidation of the basic molecule on which one, two, or three oxygen atoms can be added (Figure 8a) corresponding to mono-, di-, and trihydroxybiphenyl, respectively. Although LDI-TOF-MS analyses are considered to be mild when working slightly above the ionization threshold, it is possible that the fragment ions at m/z 91, 105, and 143 may be attributed to laser-induced rearrangements during the chemical analysis. On the other hand, we also notice the presence of larger molecules (Figure 8b) obtained by dimerization of two monohydroxybiphenyl (C24H18O2, m/z 338) and by one mono-hydroxybiphenyl and one dihydroxybiphenyl (C24H18O3, m/z 354 amu). In Figure 9, the negative ion mass spectrum is depicted for a 10 min-treatment. It conrms the presence of biphenyl
Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

molecules with 13 and even 4 added oxygen atoms. Nevertheless, intense detection of such ions could be interpreted as carboxylates, whose detection in the negative mode is enhanced compared to the positive ion mode. Fragments of biphenyl and oxidized products are also detected at m/z 88.99 and 103.01 amu. To conrm this identication, we performed LDIFTMS analyses (Figure 10). Mass resolving power reaches almost 106, giving accessed to exact mass measurement of the ions [associated mass accuracy in the (sub-)ppm range]. Results are given in Figure 10 for a 3 min-treatment. We clearly observe the presence of most of the compounds identied previously in LDI-TOF-MS measurements. Some additional peaks are detected, and can be considered as insource fragments due to a increased laser power density used in FTMS measurements compared to LDI-TOF-MS. All these ions are the overlap of contributions due to various isomers (e.g., hydroxylates or carboxylates) but LDITOF-MS and LDI-FTMS cannot separate isomers. Gas chromatography combined with mass spectrometry (GC/MS) was then used to achieve isomer separation. Results are given in Figure 11 for an untreated sample and for a sample treated for 3 min in late post-discharge (similar spectra were obtained at 1, 5, and 10 min). On the untreated sample, an intense retention peak associated with the biphenyl molecule is obtained at 7.3 min. When the sample is treated by the post-discharge, new contributions

www.plasma-polymers.org

211

W. DalMaz Silva et al.

Table 1. List of possible molecules. Isomers are not taken into account.

Detected ion

Parent moleculea)

Possible molecule

Detected ion

Parent molecule

Possible molecule
CH2

C7H 5 (88.99) C7H3O (103.01) C6H5O 3 (125.03)

C7H6 (89.039) C7H4O (103.018)


OH OH OH
O

C7 H 7 (91.05) C7H5O (105.03)

C7H7 (91.054) C7H5O (105.034)

OH

C6H6O3 (125.024)

C11H 11 (143.07)

C11H11 (143.086)

C9H7O 2 (147.04)

C9H8O2 (147.044)

OH O

C12H 10 (154.08)

C12H10 (154.078)

C12H9O (169.07)

C12H10O (169.065)

OH

C12H10O (170.07) C12H10O (170.073)

OH OH

C12H9O 2 (185.06) C12H10O2 (185.060) HO C12H9O 3 (201.06) C12H10O3 (201.055)


OH HO
OH OH HO

HO OH C12H10O 2 (186.07) C12H10O2 (186.068)

OH

C12H10O 3 (202.07) C12H10O3 (202.063)

OH HO OH
OH OH

C12H9O 4 (217.05) C12H10O4 (217.050)

OH

C24H18O 2 (338)

C24H18O2 (338.131)

OH O O OH

After chemical derivatization

C10H14SiO2 (194.076)

OH O

C24H18O 3 (354)

C24H18O3 (354.126) OH

OH OH

a)

For negative ions, the mass of the parent molecule is given with one less hydrogen atom.

appear. A rst peak is associated with the o-hydroxybiphenyl, showing a functionalization of the ring by O atoms in an alcohol group. A second series of peaks appears for retention times higher than 9 min due to isomers of dihydroxybiphenyl and other isomers of monohydroxybiphenyl. In 12, single ion monitoring performed at 169 ([M H], whose formation is favored for the ortho isomer) and 170 amu (M for the three isomers) is performed to selectively display the three isomers of the C12H10O (ortho-, meta-, and para-hydroxybiphenyl). The same kind of data treatment was performed for the C12H10O 2 fragments at
Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

186 amu gives 9 isomers among the 13 possible isomers, but the intensity of each isomer is too weak to separate them (data not shown). The analysis of these mono- and di-hydroxylated compounds by GC/MS can be performed after chemical derivatization by using combination reactive compounds: N,O-bis(trimethylsilyl) triuoroacetamide CF3CNSi(CH3)3OSi(CH3)3 (BSTFA) and trimethylchlorosilaneClSi(CH3)3 (TMCS). A detailed description of this method is given in ref.[36] The TMS derivatives are more easily separated than the hydroxylated isomers without chemical derivatization (data not shown).

212

DOI: 10.1002/ppap.201100119

Interaction Mechanisms between ArO2 Post-discharge and Biphenyl

Figure 12. Single ion monitoring chromatogram obtained at m/z 170 amu of the untreated biphenyl and the biphenyl treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa. Figure 9. Negative ion LDI-TOF-MS spectrum (from 70 to 260 amu) of the biphenyl treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa.

Benzoic acid is detected after a 10 min-treatment by GC/ MS after derivatization (Figure 13). The presence of this compound results from the opening of a ring by active species in the post-discharge.

3.3. Etching Mechanism Under the present experimental conditions used to create the post-discharge, the most concentrated species are O2 (1.0 1017 cm3), O2 (a1Dg) (1.0 1016 cm3), and O (1015 cm3).[16] Ozone was shown to be negligible[23] and other excited states of O and O2, at low concentrations

Figure 10. LDI-FTMS spectrum (from 130 to 215 amu) of the biphenyl treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa.

Figure 11. Gas chromatography combined with mass spectrometry (GC/MS) chromatograms of the untreated biphenyl and the biphenyl treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa.
Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 13. Single ion monitoring chromatogram obtained at m/z 105 179 amu, corresponding to the TMS derivatives of the benzoic acid. The biphenyl is treated for 10 min. Initial temperature: 333 K, pressure: 400 Pa. (a) mass spectrum. (b) NIST reference spectrum[37] (c) SIM chromatogram.

www.plasma-polymers.org

213

W. DalMaz Silva et al.

(<1013 cm3) will not be considered. Photon emission from the plasma and from the post-discharge is negligible.[20] We checked that the biphenyl was not modied by a simple temperature rise up to 338 2 K at 400 Pa without plasma. Therefore, O2 alone does not induce any modication in the biphenyl. Oxygen atoms do react with the biphenyl to produce the OH functions identied by the various analysis techniques that we used. Two possible pathways can be proposed. Following several authors who studied mild oxidation by catalysis,[38,39] we may assume a one-step direct insertion of atomic oxygen in the CH bond

played by the singlet state of O2.[38,39] An experiment was carried out to attempt to prove it. A 1 mm thick tungsten pipe was inserted in the 3-cm Pyrex tube to coat the walls. Recombination of oxygen atoms is known to be fast on this surface, whereas the singlet state does not deexcite on it.[41] No signicant modication in the weight loss of the sample could be measured, demonstrating the principal role of the singlet state of molecular oxygen in the etching mechanism of the biphenyl. Several works[4245] were devoted to the reactions of oxygen singlet state O2(a1Dg) with organic compounds. Among the possible reactions, one nds:[42] (i) ene reactions that occur with isolated double bonds and addition reactions, (ii) [2 2] cycloaddition that gives dioxetanes, (iii) [4 2] cycloaddition with the s-cis conformation of 1,3 diene that forms endoperoxides.

O (2p4 3P)

OH

(b)

Certainly, this process occurs several times leading to di-, tri-, and tetra-hydroxybiphenyl. Such a mechanism cannot explain the presence of molecular chains longer than the biphenyl. The presence of reticulated chains shows that cross-linking reactions occur. Consequently, radicals must be formed from biphenyl molecules. For olens, formation of a radical is explained by H abstraction with an oxygen atom, contrary to the previous mechanism. Therefore, we assume that another pathway prevails

ene reactions are possible with olens[45] that could be present once a ring is opened, i.e., as by-products of the biphenyl oxidation. It happens when allylic hydrogen are oxidized to allylic hydroperoxides concomitant with a shift of the double bond

O (2p4 3P) +

H O2(a1g) + H

ene reaction H


(e)

+ OH

(c)

This second pathway occurs either simultaneously to the rst one, or when the temperature changes. From then on, reticulation results from the following process

+H

Because of the relative weakness of the peroxide bond, its homolysis to alkoxy radicals at room temperature or above is a prevalent phenomenon.[45] Three mechanisms are possible but two of them require atomic hydrogen and we shall see that there is a decit of H atoms. The last one is b cleavage of a neighboring b hydrogen

(d)

H
As the tetraphenyl molecule is detected with OH groups grafted on it, we may think that process (d) and then process (c) start only once process (b) has already occurred, i.e., at higher temperature, and not simultaneously. Oxidative coupling of aromatic compounds with molecular oxygen could also be mentioned to explain crosslinking reactions if we assume that traces of metal catalysts are present in the biphenyl, but much higher temperature and molecular oxygen pressure than those used in this work are usually needed.[40] We shall not consider that oxygen atoms could be responsible for the ring opening. Most of the results on mild oxidation of aromatic compounds rather stress the role
Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

H H + ROH
(f)

Addition reactions are also possible, leading to the formation of ketones


Addition O O
g)

O2

(a1

+ O+O

(g)

A second set of reactions deals with [2 2] cycloaddition. This process has been observed on carbon nanotubes[43] for instance.

214

DOI: 10.1002/ppap.201100119

Interaction Mechanisms between ArO2 Post-discharge and Biphenyl

[2+2] cycloaddition O2 (a1


g)

O O
(h)

containing compound, it can only be the maleic anhydride. We can resort to a [4 2] cycloaddition to propose the following mechanism
[4+2] cycloaddition OH
O O

In general, dioxetanes cleave thermally or photochemically producing carbonyl fragment and chemiluminence[44]
O O O O O

O2(a1g) + OH O2(a1g) + OH

OH OH
O O

(m)
O +O OH O O

O O

O O

O O

(i)

+ O O

O O O O

This process is possibly the rst step of the biphenyl etching, leading to the opening of one of the rings. However, these mechanisms produce carbonyl compounds that could be present but are not detected in our analyses. Indeed, for compounds with large masses, like C12H10O4 (see Table 1) it is possible to assume the presence of carbonyl groups, but their presence is not demonstrated. On the contrary, alcohols or epoxides are detected (C7H5O and C7H4O for instance). The detection of the tropylium ion (C7H 7 ) is justied by the high-stability of the benzyl radicals[46] and by the presence of C7H6 that results from its thermal decomposition[47]

The melting point of maleic anhydride is 326 K. It is almost exactly the minimum temperature needed to etch the biphenyl (Figure 5). We observed experimentally a white deposit of a compound downstream the crucible where the biphenyl is located, in a colder part of the reactor. Due to the weak amount of product condensed on the walls, we could not analyze it but we may reasonably assume it is the maleic anhydride. Let us notice that in the case of the biphenyl, rotation about the single bond is known to be sterically hindered. Consequently, we can reasonably assume that there is no [4 2] cycloaddition between the singlet state and the biphenyl over the CC bond in between the two rings

HO O H O2(a1g) + endoperoxide
(n)

4. Conclusion
We can draw the following sketch presented in Figure 14. Oxygen atoms are supposed to create OH functional groups on the biphenyl and to produce radicals which can undergo reticulation. The oxygen singlet state may react according to various chemical pathways but [4 2] cycloaddition

CH2 +H
(j)

Consequently, the presence of benzoic acid can result from the following reaction described by Yoshino et al.[48]

OH O2 + CH2 +H O
(k)

According to,[49] the benzyl radicals can be synthesized by the following reaction
O CH2 + CH2O

(l)

This emphasizes a ring opening rather based on an alkoxy compound than on the cleavage of dioxetanes. The origin of the radical of the 2-phenethyl alcohol is the biphenyl oxidation. The other fragment being necessarily a C4HPlasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 14. Possible reaction pathways leading to the by-products identied with the analysis techniques used in this work.

www.plasma-polymers.org

215

W. DalMaz Silva et al.

gives by-products which are detected by the our analysis techniques. Finally, the activation energy of the etching process is coherent with values encountered for thermolyses of endoperoxides.
Acknowledgements: The authors thank L. Vernex-Loset and G. Krier for the access to the MS instruments at the University of Metz (France) and for their scientic expertise. They also thank the CAPES/COFECUB, a joint Brazilian/French project (Ph 697/10), for support. These Researches were carried out within the framework of the Associated European Laboratory: Laboratoire dInteraction Plasma Extre me Surface (LIPES).

Received: June 12, 2011; Revised: September 14, 2011; Accepted: September 21, 2011; DOI: 10.1002/ppap.201100119 Keywords: biphenyl; lms; melting point; plasma cleaning; postdischarge

[1] C. D. Pintassilgo, K. Kutasi, J. Loureiro, Plasma Sources Sci. Technol. 2007, 16, S115. [2] M. J. Pinheiro, G. Gousset, A. Granier, C. M. Ferreira, Plasma Sources Sci. Technol. 1998, 7, 524 & PSST 8 31 (1999). [3] K. Kutasi, B. Saoudi, C. D. Pintassilgo, J. Loureiro, M. Moisan, Plasma Process. Polym. 2008, 5, 840. [4] K. Kutasi, V. Guerra, P. Sa, J. Phys. D: Appl. Phys. 2010, 43, 175201. [5] V. Guerra, K. Kutasi, P. A. Sa, Appl. Phys. Lett. 2010, 96, 071503. [6] K. Kutasi, V. Guerra, P. A. Sa , J. Loureiro, Active species in ArO2 and ArO2N2 owing microwave discharges and post-discharges, Proc. of 19th International Symposium on Plasma Chemistry Bochum, Germany 2009. [7] G. Primc, R. Zaplotnik, A. Vesel, M. Mozetic, AIP Adv. 2011, 1, 022129. [8] C. Canal, F. Gaboriau, A. Ricard, M. Mozetic, U. Cvelbar, A. Drenik, Plasma Chem. Plasma Process. 2007, 27, 404. [9] A. Ricard, V. Monna, M. Mozetic, Surf. Coat. Technol. 2003, 174, 905. , J. Phys.: Conf. Ser. 2009, 162, 012015. [10] A. Vesel, M. Mozetic , [11] I. Junkar, U. Cvelbar, A. Vesel, N. Hauptman, M. Mozetic Plasma Process. Polym. 2009, 6, 667. , S. Strnad, Vacuum [12] I. Junkar, U. Cvelbar, A. Vesel, M. Mozetic 2009, 84, 83. [13] H. Yasuda, M. Gazicki, Biomaterials 1982, 3, 68. [14] J. Larrieu, B. Held, F. Cle ment, R. C. Hiorns, Eur. Phys. J. Appl. Phys. 2003, 22, 61. [15] M. Laroussi, Plasma Process. Polym. 2005, 2, 391. [16] G. Fridman, G. Friedman, A. Gutsol, A. B. Shekhter, V. N. Vasilets, A. Fridman, Plasma Process. Polym. 2008, 5, 503. [17] M. Moisan, J. Barbeau, M.-C. Crevier, J. Pelletier, N. Philip, B. Saoudi, Pure Appl. Chem. 2002, 74, 349. [18] S. Lerouge, M. R. Wertheimer, LH. Yahia, Plasmas Polym. 2001, 6, 175. [19] L. Lefe ` vre, T. Belmonte, T. Czerwiec, A. Ricard, H. Michel, Appl. Surf. Sci. 2000, 153, 85. [20] V. Hody, T. Belmonte, C. Pintassilgo, F. Poncin-Epaillard, T. Czerwiec, G. Henrion, Y. Segui, J. Loureiro, Plasma Chem. Plasma Proc. 2006, 26, 251.

[21] V. Hody, T. Belmonte, T. Czerwiec, G. Henrion, J. M. Thie baut, Thin Solid Films 2006, 506507, 212. [22] T. Belmonte, C. Pintassilgo, T. Czerwiec, G. Henrion, V. Hody, J. M. Thie baut, J. Loureiro, Surf. Coat. Technol. 2005, 200, 26. [23] M. Mafra, T. Belmonte, F. Poncin-Epaillard, A. S. da Silva Sobrinho, A. Maliska, Plasma Chem. Plasma Process. 2008, 28, 495. [24] M. Mafra, T. Belmonte, A. Maliska, A. S. da Silva Sobrinho, U. Cvelbar, F. Poncin-Epaillard, Key Eng. Mater. 2008, 373374, 421. [25] C. Noe l, D. Duday, S. Verdier, P. Choquet, T. Belmonte, H.-N. Migeon, Plasma Process. Polym. 2009, 6, S187. [26] E. Bernardelli, T. Belmonte, D. Duday, G. Frache, F. PoncinEpaillard, A. Maliska, Plasma Chem. Plasma Process. 2011, 31, 189. [27] E. Bernardelli, T. Belmonte, D. Duday, G. Frache, F. PoncinEpaillard, A. Maliska, Plasma Chem. Plasma Process. 2011, 31, 205. [28] E. Bernardelli, A. Ricard, T. Belmonte, Plasma Sources Sci. Technol. 2011, 20, 025012 [29] E. Gonzalez, M. D. Barankin, P. C. Guschl, R. F. Hicks, IEEE Trans. Plasma Sci. 2009, 37, 823. [30] F. D. Eggito, Pure Appl. Chem. 1990, 62, 1699. [31] F. D. Eggito, L. J. Matienzo, IBM J. Res. Dev. 1994, 38, 423. [32] F. D. Eggito, F. Emmi, R. S. Horwath, V. Vukanovic, J. Vac. Sci. Technol., B 1985, 3, 893. [33] L. Lefe ` vre, T. Belmonte, H. Michel, J. Appl. Phys. 2000, 8710, 7497. [34] N. J. Two, M.-F. Chow, J. Rigaudy, J. Am. Chem. Soc. 1981, 103, 7218. [35] G. M. Brown, B. S. Brunschwig, C. Creutz, J. F. Endicott, N. Sutin, J. Am. Chem. Soc. 1979, 101, 1300. [36] D. Li, J. Park, J.-R. Oh, Anal. Chem. 2001, 73, 3089. [37] Mass Spectra by NIST Mass Spec Data Center, S.E. Stein, director. NIST chemistry WebBook, Standard Reference Database number 69, Eds. P.J. Linstrom and W.G. Mallard, National Institute of Standards and Technology, Gaithersburg MD, 20899, http://webbook.nist.gov. [38] K. Teramura, T. Tanaka, T. Hosokawa, T. Ohuchi, M. Kani, T. Funabiki, Catal. Today 2004, 96, 205. [39] R. L. Cargill, Benjamin M. Gimarc, David M. Pond, Thomas Y. King, A. Bradford Sears, M. Robert Willcott, J. Am. Ceram. Soc. 1970, 92, 3810. [40] R. D. Vidic, M. T. Suldan, R. C. Brenner, Environ. Sci. Technol. 1993, 27, 2079. [41] F. Normand, J. Marec, Ph. Leprince, A. Granier, Mater. Sci. Eng., A 1991, 139, 103. [42] C. S. Foote, Ed., Active Oxygen in Chemistry, Chapman & Hall, London 1995, p. 114. [43] S.-P. Chan, G. Chen, X. G. Gong, Z.-F. Liu, Phys. Rev. Lett. 2003, 90, 086403. [44] A. A. Frimer, Singlet Oxygen in Peroxide Chemistry, in The Chemistry of Functional Groups, Peroxides, S. Patai, Ed., John Wiley & Sons Ltd, 1983, pp. 201234. [45] A. A. Frimer, Chem. Rev. 1979, 79, 359. [46] D. C. Nonhebel, J. C. Walton, Free Radical Chemistry, Cambridge University Press, 1974. [47] G. da Silva, J. A. Cole, J. W. Bozzelli, J. Phys. Chem. A. 2009, 113, 6111. [48] Y. Yoshino, Y. Hayashi, T. Iwahama, S. Sakaguchi, Y. Ishii, J. Org. Chem. 1997, 62, 6810. [49] F. P. Lossing, Can. J. Chem. 1971, 49, 357.

216

Plasma Process. Polym. 2012, 9, 207216 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

DOI: 10.1002/ppap.201100119

Potrebbero piacerti anche