44
1 Volatile products controlling Titan’s tholins production 1 Nathalie Carrasco 1 *, Thomas Gautier 1 , Et-touhami Es-sebbar 1 , Pascal Pernot 2 , Guy 2 Cernogora 1 3 4 1 Laboratoire Atmosphères, Milieux, Observations Spatiales, Université de Versailles Saint- 5 Quentin, UMR 8190, 78280 Guyancourt, France 6 2 Laboratoire de Chimie Physique, UMR 8000, CNRS, Univ Paris-Sud, 91405 Orsay cedex, 7 France 8 * Electronic Address: [email protected] 9 Abstract 10 A quantitative agreement between nitrile relative abundances and Titan’s atmospheric 11 composition was recently shown with a reactor simulating the global chemistry occurring in 12 Titan’s atmosphere [Gautier et al. (2011) Icarus, 213: 625]. Here we present a complementary 13 study on the same reactor using an in-situ diagnostic of the gas phase composition. Various 14 initial N 2 -CH 4 gas mixtures (methane varying from 1 to 10 %) are studied, with a monitoring 15 of the methane consumption and of the stable gas neutrals by in-situ mass spectrometry. 16 Atomic hydrogen is also measured by optical emission spectroscopy. A positive correlation is 17 found between atomic hydrogen abundance and the inhibition function for aerosol production. 18 This confirms the suspected role of hydrogen as an inhibitor of heterogeneous organic growth 19 processes, as found in [Sciamma-O’Brien et al. (2010) Icarus, 209, 704]. The study of the gas 20 phase organic products is focussed on its evolution with the initial methane amount [CH 4 ] 0 21 and its comparison with the aerosol production efficiency. We identify a change in the 22 stationary gas phase composition for intermediate methane amounts: below [CH 4 ] 0 =5%, the 23 gas phase composition is mainly dominated by Nitrogen-containing species, whereas 24

Volatile products controlling Titan’s tholins production

  • Upload
    cearc

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

1

Volatile products controlling Titan’s tholins produ ction 1

Nathalie Carrasco1*, Thomas Gautier1, Et-touhami Es-sebbar1, Pascal Pernot2, Guy 2

Cernogora1 3

4

1Laboratoire Atmosphères, Milieux, Observations Spatiales, Université de Versailles Saint-5

Quentin, UMR 8190, 78280 Guyancourt, France 6

2Laboratoire de Chimie Physique, UMR 8000, CNRS, Univ Paris-Sud, 91405 Orsay cedex, 7

France 8

* Electronic Address: [email protected] 9

Abstract 10

A quantitative agreement between nitrile relative abundances and Titan’s atmospheric 11

composition was recently shown with a reactor simulating the global chemistry occurring in 12

Titan’s atmosphere [Gautier et al. (2011) Icarus, 213: 625]. Here we present a complementary 13

study on the same reactor using an in-situ diagnostic of the gas phase composition. Various 14

initial N2-CH4 gas mixtures (methane varying from 1 to 10 %) are studied, with a monitoring 15

of the methane consumption and of the stable gas neutrals by in-situ mass spectrometry. 16

Atomic hydrogen is also measured by optical emission spectroscopy. A positive correlation is 17

found between atomic hydrogen abundance and the inhibition function for aerosol production. 18

This confirms the suspected role of hydrogen as an inhibitor of heterogeneous organic growth 19

processes, as found in [Sciamma-O’Brien et al. (2010) Icarus, 209, 704]. The study of the gas 20

phase organic products is focussed on its evolution with the initial methane amount [CH4]0 21

and its comparison with the aerosol production efficiency. We identify a change in the 22

stationary gas phase composition for intermediate methane amounts: below [CH4]0=5%, the 23

gas phase composition is mainly dominated by Nitrogen-containing species, whereas 24

2

hydrocarbons are massively produced for [CH4]0>5%. This predominance of N-containing 25

species at lower initial methane amount, compared with the maximum gas-to solid conversion 26

observed in Sciamma-O’Brien et al. 2010 for identical methane amounts confirms the central 27

role played by N-containing gas-phase compounds to produce tholins. Moreover, two 28

protonated imines (methanimine CH2=NH and ethanamine CH3CH2=NH) are detected in the 29

ion composition in agreement with Titan’s INMS measurements, and reinforcing the 30

suspected role of these chemical species on aerosol production. 31

32

3

1-Introduction 33

On Titan, the dissociation of N2 and CH4 by solar UV radiation, cosmic rays and Saturn’s 34

magnetosphere electron bombardment induces a complex organic chemistry at high altitudes 35

(Waite Jr. et al., 2007) that results in the production of solid aerosols responsible for the 36

orange haze surrounding Titan. These are the most complex extraterrestrial organic material 37

detected in the solar system. Their chemical production mechanisms are roughly described, 38

first by a photolytic priming of the nitrogen and methane precursors, then by the formation of 39

hydrocarbon and nitrile monomers in the gaseous phase, finally followed by polymerization 40

processes leading after recombination to solid particles big enough to condensate, 41

agglomerate and settle down to the ground. However, if one looks in the detail, each step 42

identified in the chemical growth process is subject to research and to continuous 43

developments. The work of Imanaka et Smith, 2007 showed the central role of N2 ionization 44

for gas products formation. Pernot et al., 2010 suggested new routes towards polymerization, 45

complementary to the previous acetylene, benzene, HCN and HC3N routes implemented in 46

Titan's photochemical models (Lebonnois, 2005; Lebonnois et al., 2002). 47

Several experimental setups have been developed in order to reproduce and study in the 48

laboratory such a complex atmospheric system. Among them, the plasma device PAMPRE 49

provided significant clues on the understanding of the polymeric chemical structure of the 50

aerosols (Carrasco et al., 2009; Pernot et al., 2010). The influence of the methane initial 51

concentration on the aerosol mass production efficiency was moreover studied in Sciamma-52

O'Brien et al., 2010. A maximum has been found for the intermediate initial concentrations of 53

methane. This important result highlights a competition between a polymerization growth 54

process correlated with the methane concentration, and an inhibition process anti-correlated 55

with the methane concentration. Several hypotheses were proposed to explain this inhibition 56

process in the aerosol production, all of them involving the increase of the global hydrogen 57

4

content in the gas phase coming from methane dissociation: either by saturation of the 58

growing solid grains with hydrogen (molecular or atomic), or by production in the gas phase 59

of saturated volatiles less favourable to achieve polymerization. In order to find some clues on 60

this sensitivity of the aerosol production with the methane initial content in the gas mixture, 61

we began to study the evolution of the gas phase products composition according to this 62

initial methane amount. The stable neutrals were detected by in-situ mass spectrometry. 63

Moreover a relative quantification of the atomic hydrogen content was performed by in-situ 64

Optical Emission Spectroscopy (OES). All the in-situ products found in this work are to be 65

compared with the parallel study of Gautier et al., 2011 on the same PAMPRE reactive gas 66

mixtures analysed by cryogenic trapping and Gas Chromatography-Mass Spectrometry (GC-67

MS) and with the previous gas mixture analysis on similar plasma experiments (Coll et al., 68

1999; De Vanssay E et Raulin, 1995; Ramírez et al., 2001; Ramírez et al., 2005). 69

70

5

2-Experimental 71

2.1. The PAMPRE plasma reactor 72

The PAMPRE experimental setup has been described in detail in previous publications 73

(Alcouffe et al., 2010; Szopa et al., 2006). Briefly, the reactor consists of a cylindrical 74

stainless steel chamber. Two gas bottles, one of pure N2 and one containing a N2-CH4 mixture 75

at 10% CH4 concentration, are used to obtain different gas mixtures in the experiment. A third 76

bottle of Argon is used to introduce Argon as an actinometer (see Part 2.3.). The Radio 77

Frequency Capacitively Coupled Plasma (RF-CCP) discharge is driven by a 13.56 MHz 78

frequency generator with power up to 100 W, and generated between two electrodes: one 79

grounded, one polarized. The discharge is confined by the grounded electrode which is a 80

cylindrical grid box, 13.7 cm in diameter and 4 to 5 cm in height. The N2-CH4 gas mixture is 81

continuously injected through the polarized electrode and pumped by a rotary vane vacuum 82

pump. With this discharge design, the gas flow is uniform in the confining cylindrical box. 83

In the study presented here, the RF power is fixed at 30 W, the total N2-CH4 gas flow rate at 84

55 sccm and the gas pressure at 0.9 mbar for all the experiments. A constant argon gas flow 85

rate of 2.75 sccm (corresponding to an amount of ~5%) is added to the gas mixture for 86

actinometry measurements. This compound does not participate to the gas phase chemistry, 87

and its low amount does not modify the plasma discharge. 88

The low pressure RF discharge produces a “cold plasma”, which is not at thermodynamical 89

equilibrium. The neutral gas remains at ambient temperature (Alcouffe et al., 2010), when 90

electron have an mean energy on the order of one or two eV. The assumed maxwellian 91

electron energy distribution mimics quite correctly the photon solar energy distribution 92

(Szopa et al., 2006). Neutrals are warmer in the PAMPRE experiment than in Titan’s 93

6

atmospheric conditions, globally enhancing the kinetics of neutral-neutral bimolecular 94

reactions. 95

Before each experiment, the reactor chamber, plasma box and crystallizer are cleaned with 96

ethanol. The reactor is then pumped down to secondary vacuum by a turbo molecular pump 97

down to 2×10-6 mbar, and baked out at 110°C for several hours in order to remove impurities 98

adsorbed onto the chamber walls. Once the chamber has cooled down, a 30 minutes argon 99

plasma ensures the outgassing of the metallic confining cylindrical grid box. 100

2.2. Mass spectrometry 101

In situ measurements of the gas phase composition are achieved using a Pfeiffer QME 200 102

quadrupole mass spectrometer (MS). In the MS, the neutral molecules are ionized by electron 103

impact at the standard 70 eV electron energy. The gas sampling is done through a capillary 104

tube (0.8 mm in diameter), with a sampling orifice located 1 cm outside the reactive plasma. 105

The capillary is long enough to reduce the gas flow rate between the reactor chamber and the 106

MS and keep the operating pressure in the MS below 10-5 mbar. As a consequence, only 107

stable molecules can be measured with this experimental set up. The MS detector has a 108

resolution of 1 u and covers the 1 to 100 u range. 109

A weak air signature (N2, O2, CO2) is highlighted in the mass spectra leading to contributions 110

at m/z 14, 16, 28, 32 and 44, even in the blank of the mass spectrometer given Figure 1. The 111

blank corresponds to the isolated mass spectrometer functioning at a vacuum limit of about 112

3×10-8 mbar. The water signal at m/z= 2, 16, 17, 18 is quite important, reaching an intensity 113

of about 1×10-10 A for m/z=18 in all the spectra including the blank and remained stable 114

during the experiments. This signal is due to remaining water adsorbed on the wall of the 115

QMS housing because a very high temperature outgassing was not possible with the used 116

configuration. Figure 1 also shows a spectrum recorded for a 5% methane concentration in the 117

PAMPRE reactor when the plasma is OFF. The m/z = 18 signal remains the same. An 118

7

intensity threshold for recording has been chosen equal to 10-13 A. Studying the stationary 119

state allows long acquisition time-scales: a 2 seconds shift is chosen for each mass between 120

m/z 2 to 80. 121

122

1.00E-14

1.00E-13

1.00E-12

1.00E-11

1.00E-10

1.00E-09

1.00E-08

0 5 10 15 20 25 30 35 40 45 50 55 60

m/z

Inte

nsity

(A

)

123

Figure 1: (Black) Blank mass spectrum of the mass spectrometer only. The low residual air signal is 124

consistent with the vacuum limit pressure of 3×10-8 mbar. The quite important water signal, at m/z=18, 17, 125

16 and 2 is stable, and corresponds to residual water adsorbed in the QMS. (Red) Initial mass spectrum 126

obtained with a N2-CH4 gas mixture of [CH4]0=5%. 127

128

The methods used are systematically both scan analog (range of masses continuously scanned 129

and recorded) and scan bargraph mode (only intensities of unit mass peaks are recorded). The 130

bargraph mode is faster and allows a direct comparison with the NIST database (Stein, 2011), 131

but we observed that this acquisition mode can miss some products with mass peak in the tail 132

of an adjacent predominant species. An example is given on Figure 2 showing the respective 133

analog and bargraph acquisitions for the same gas mixture obtained after a plasma discharge 134

8

in a 1% methane gas mixture. The HCN production (m/z 27) is visible in the analog scan, but 135

has been completely missed in the bargraph mode because of the major N2+ fragment (m/z 136

28). Both acquisitions were thus systematically recorded and cross-compared, helping to 137

correct the possible incomplete bargraphs. 138

139

1,0E-14

1,0E-13

1,0E-12

1,0E-11

1,0E-10

1,0E-09

1,0E-08

0 5 10 15 20 25 30 35 40 45 50 55 60

m/z

Inte

nsity

(A

)

140

Figure 2: Comparison between a scan analog (red line) and a scan bargraph acquisition (black squares) 141

on the example of the stationary state plasma discharge in a 1% methane initial gas mixture. 142

143

2.3. Optical emission spectroscopy and actinometry 144

The concentration of atomic hydrogen cannot be studied with our mass spectrometry device, 145

which measures only stable neutrals. Optical Emission Spectroscopy (OES) is a non-intrusive 146

diagnostic method commonly used for in situ studies of plasma discharges. In our conditions, 147

the H atoms produced from CH4 dissociation are excited by electronic collision to radiative 148

levels. These levels are then depopulated by radiative emission. The intensity of an H line is 149

given by: 150

9

( )[ ]HTknHA

HAHhCVHI eHee

ji

jiHO −

∑=

)(

)()()()(

,

,νλ (1) 151

• where VO is the volume of the plasma observed by OES, 152

• C(λH) is a spectral calibration factor of the optical device for the chosen hydrogen line 153

wavelength, 154

• hν(H) is the photon energy, 155

• A i,j the Einstein coefficient for the observed line, 156

• ∑ )(, HA ji =1/τ (τ is the radiative life time of the level). 157

These five terms are independent on the plasma and on the H concentration. 158

The three last are plasma depending: ne is the electron density, ke-H(Te) the rate coefficient for 159

the population of the radiative level, function of the electron temperature Te and [H] is the 160

atomic hydrogen density. 161

Equation (1) shows that even if an H line intensity is proportional to [H], it depends also on 162

plasma parameters. 163

In order to avoid this problem, we add into the plasma a small amount of argon as 164

actinometer. Argon is non-reactive and, in our experimental conditions, its radiative levels are 165

populated only by electron collisions and depopulated by spontaneous emission as H lines. 166

Then, an Ar line intensity is given by: 167

( )[ ]ArTknArA

ArAArhCVArI eAree

ji

jiArO −

∑=

)(

)()()()(

,

,νλ (2) 168

where VO is the same observed volume as for the H line and C(λAr) is the spectral calibration 169

factor at the Ar line wavelength. The Ai,j are the Einstein coefficients for the Ar line. 170

Combining equations (1) and (2) the H atoms density is given by: 171

[ ] ( )( )

( )( ) [ ]ArTk

Tk

ArI

HIAH

eHe

eAre

−= (3) 172

10

As mentioned before, the Ar amount is constant. So, equation (3) shows that [H] is 173

proportional to the ratio of the intensity of select H and Ar lines and not to electron density: it 174

is the optical actinometry method. The question of the electronic rate coefficients will be 175

discussed below. 176

The A factor depends only on spectroscopic data of selected lines and spectral calibration, but 177

not on the observed volume. 178

Before using optical actinometry, we have to check if the addition of Ar in the gas mixture 179

modifies the plasma discharge. As explained in Alcouffe et al. (2010) the self-bias voltage Vdc 180

of the polarized electrode is directly correlated with electron density and temperature. No 181

variation of Vdc has been measured when argon is added to the discharge. 182

For OES, the light emitted by the plasma is dispersed by a 60 cm focal length monochromator 183

(Jobin-Yvon) connected to a photomultiplier (Hamamatsu R928). The output current of the 184

photomultiplier is measured with a picoammeter (Keithley 6485) connected to a computer 185

(Alcouffe et al., 2010). Spectra are recorded in the visible to the near IR using a colored glass 186

filter to eliminate overlapping of second order lines in the red and near IR range. 187

A spectrum of the discharge from 475 nm to 825 nm is presented in Figure 3. The spectrum is 188

clearly dominated by nitrogen molecular bands emission of second positive system (SPS) 189

N2(C→ B) and first positive system (FPS) N2(B→A), hydrogen emission of Hβ (486.1 nm) 190

and Hα (656.3 nm), and atomic Ar lines . 191

192

11

475 480 485 490

0

10

20

30

640 645 650 655 6600

50

100

150

740 745 750 755 7600

200

400

600

810 815 820 8250

50

100

150

*

N2(B

,A)(

9,6)

N2(B

,A)(

8,5)

Ar,

826

.9nm

Ar,

811

.5nm

Ar,

810

.4nm

N2(B

,A)(

7,4)

Ar,

751

.5nm

Ar,

750

.4nm

H ββ ββ

N2(C

,B)(

2,8)

Inte

nsity

, a.u H αα αα

5 % CH4/ 5 % Ar/ N

2

Wavelength, nm 193

Figure 3: Typical example of emission spectrum obtained between 475 and 828 nm for 5% CH4/N2 RF 194

plasma with the presence of 5% Ar as an actinometer gas. The mean features emissions are identified 195

together their wavelengths. The asterisk points out the N2(B, A)(6, 3) emission line. 196

197

Gicquel et al., 1998 have used optical actinometry with Ar to determine H concentration in 198

H2-CH4 plasma. In order to have limited variations of the electronic rate constants in the 199

equation (3), Gicquel et al. use the Hα (656.3 nm) and the 750.3 nm Ar lines whose excitation 200

energies are on the same order of magnitude (respectively 12.1 eV and 13.5 eV), and the cross 201

sections for excitation by electron collisions are similar. 202

But in our case, Figure 3 shows that in N2-CH4 plasma the Ar (750.4 nm) line is 203

superimposed to the head band of the (4-2) and the Hα line is embedded in the rotational 204

spectra of the (6-3) band of the FPS. On the opposite, the Hβ (486.1 nm) as Ar (811.5 nm) 205

lines are well isolated. As their excitation energies, 12.74 eV for Hβ and 13.07 eV for Ar 206

12

(811.5 nm), are on the same order of magnitude and in agreement with the criteria defined by 207

Gicquel et al., 1998, these two lines are chosen for optical actinometry measurements. 208

209

13

3-Results and discussion 210

Various initial gas mixtures are studied, with 1, 2, 5, 8 and 10% of methane in nitrogen, 211

covering the methane concentration range of our previous study on the gas-to solid 212

conversion yield (Sciamma-O'Brien et al., 2010). 213

3.1. Atomic hydrogen 214

Optical Emission Spectroscopy measurements have been done for plasma conditions with 215

initial amount of CH4 from 1 to 10% in nitrogen. The argon amount was always the same 216

(about 5%). Figure 4 shows the intensity of Hβ (486.1 nm) and Ar (811.5 nm) lines as a 217

function of the CH4 percentage. 218

0 2 4 6 8 10

50

100

150

200

b: Ar(811.5nm)

Inte

nsity

, a.u

% CH4

0 2 4 6 8 103

6

9

12

Inte

nsity

, a.u

a: Hββββ(486.5nm)

219

Figure 4: Dependence of (a) Hβ (486.1 nm) and (b) Ar (811.5 nm) emission line intensities on the 220

percentage of CH4 added to N2 RF plasma. The percentage of Ar as an actinometer gas is 5% for all 221

measurements. 222

223

14

The Hβ emission reaches a maximum for about 5% of CH4. Whereas the argon density is 224

constant, Ar line varies with the CH4 amount. As shown in equation (2), this variation is 225

related to the electron density and temperature. In order to remove the influence of electron 226

density, the ratio Hβ/Ar is calculated and presented in Figure 5. 227

Equation (3) shows that this ratio is proportional to H density, [H]. 228

229

Figure 5: Evolution of hydrogen atom densities as a function of % CH 4 obtained from Figure 4 in the 230

same experimental conditions. 231

232

From Figure 5, it can be deduced that [H] increases linearly with the injected amount of CH4. 233

The linearity is valid only if the ratio ke-Ar(Te)/ke -H(Te) remains constant. Alcouffe et al., 2010 234

have shown that the electron temperature varies when solid particles are present in the plasma. 235

Gicquel et al., 1998 have calculated that this ratio variation is only about 10% when electron 236

temperature changes from 1eV to 5eV. As in our experimental conditions electron energy 237

15

remains on the order of few eV, we can consider that the ratio is constant and thus that [H] 238

increases linearly with the injected amount of CH4. 239

This positive linear increase with [CH4]0 is anti-correlated with the decrease of the gas-to-240

solid conversion efficiency measured in Sciamma-O'Brien et al., 2010. As a consequence, 241

atomic hydrogen provides a possible important clue to understand inhibition of the aerosol 242

production observed when the amount of methane increases. 243

244

3.2. Stable neutrals. 245

Before studying the reactive plasma, some general comments can be made on the reference 246

gas mixture mass spectra before plasma ignition (see Figure 6). The initial mixture is mainly 247

composed by molecular nitrogen and methane, which can be identified on the mass spectrum 248

with their following fragments. Molecular nitrogen involves fragments at m/z = 7 (N2+), 14 249

(N+), 15 (15N+) 28 (N2+), 29 (15N14N+) and 30 (15N15N+). Methane exhibits fragments at m/z 250

=12, 13, 14, 15 and 16, respectively C+, CH+, CH2+, CH3

+ and CH4+. Additional argon is 251

introduced for complementary optical emission spectroscopy measurements, leading to large 252

contributions at m/z = 20, 36 and 40. 253

254

16

0 5 10 15 20 25 30 35 40 45 50 55 6010-14

10-13

10-12

10-11

10-10

10-9

10-8

Inte

nsity

, a.u

m/z

Plasma OFF Plasma ON

255

Figure 6: Mass spectrum of a CH4-N2 10-90% gas mixture, (black) plasma OFF, (red) plasma ON. 256

257

As the absolute intensity measured by the mass spectrometer is not faithfully reproducible, 258

mainly due to the progressive clogging of the filament in the ionization chamber and a slow 259

drift of the amplification factor of the channeltron, we chose to work with relative intensities 260

among the products. For each experiment series, mass spectra are recorded before the plasma 261

discharge. The initial state is hereafter called initial condition and labelled with the subscript 262

0. When the plasma is on, mass spectra are recorded after the transient regime when the 263

discharge is at the stationary state (Sciamma-O'Brien et al., 2010). This stationary state is 264

labelled with the subscript SS. The spectra are therefore normalized with a peculiar correction 265

factor [N2]0/I0(28) where I0(28) is the intensity current (A) of the mass peak m/z 28 and [N2]0 266

the nitrogen concentrations in the initial gas mixture (plasma off). When normalized, spectra 267

are expressed in arbitrary units (AU). 268

17

3.2.1. Molecular Hydrogen 269

Information of molecular hydrogen production is contained in the m/z = 2 intensity 270

production Iss(2)-I0(2), difference between stationary state and initial mass spectra. Indeed this 271

indicator informs us about the global hydrogen content in the reactive plasma, summing the 272

major contribution of molecular hydrogen, but also of atomic hydrogen efficiently 273

recombined on the metallic sampling line before the ionization chamber of the mass 274

spectrometer. 275

Secondary contributions from hydrocarbon products fragmentation are negligible. Indeed, as 276

shown on Figure 6, the m/z=2 signal increases from 8×10-11A to 1×10-9A. A similar 277

production of a few 1×10-9A can be observed for the peak at m/z=28. It corresponds to the 278

ethane C2H4 signature, which has no significant contribution at m/z=2 in its fragmentation 279

pattern (NIST and Pfeiffer databases). All the other products peaks are at the most of the 280

order of 1×10-11A. The fragments of these products can therefore not contribute significantly 281

to the m/z=2 signal. 282

The contribution of atomic hydrogen can also be considered as negligible if we compare with 283

the concentrations of H2 and H predicted by Pintassilgo et Loureiro, 2010 for a N2-CH4 284

afterglow plasma model, corresponding to ten times less H atoms than H2 molecules. As a 285

consequence, Iss(2)-I0(2) can at first order be assimilated to molecular hydrogen signature. 286

Figure 7 reports the molecular Hydrogen production (difference between the stationary Iss(2) 287

and the initial I0(2) intensities of the mass peak at m/z=2), as a function on the initial methane 288

concentration (upper panel), or as a function of the consumed methane (lower panel). Error 289

bars are calculated as ±2σ on the statistical among several recorded spectra for each condition. 290

The uncertainty is larger for the [CH4]0=8% conditions because of an unusual drift of the 291

signal during this specific experiment. 292

293

18

0 2 4 6 8 100.0

0.1

0.2

0.3

0.4

0.5

I ss

(2)-

I 0(2)

(AU

)

[CH4]0 (%)

-0.05 -0.04 -0.03 -0.02 -0.010.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

I SS

(2)-

I 0(2

) (A

U)

ISS

(15)-I0(15) (AU)

294

Figure 7: Hydrogen production (difference between the stationary Iss(2) and the initial I0(2) intensities of 295

the mass peak at m/z=2) (above) according to the initial methane concentration in the gas mixture; (below) 296

according to the methane consumption (difference between the stationary Iss(15) and the initial I0(15) 297

intensities of the mass peak at m/z=15). The methane fragment CH3+ at m/z = 15 has been chosen because 298

it does not overlap with water fragments. 299

19

We observe on Figure 7 (upper panel) that Iss(2)-I0(2) increases globally with the initial 300

methane content, but reaches an asymptote for initial concentrations higher than 5% of 301

methane. This can be related to the same asymptote observed in Sciamma-O'Brien et al., 2010 302

for methane consumption Iss(15)-I0(15). Both parameters are strongly correlated as can be 303

seen on the lower panel of Figure 7. This H2 enrichment with the initial concentration of 304

methane is consistent with the time-dependent measurement carried out by Majumdar et al., 305

2005 with a 2:1 CH4-N2 gas mixture. Moreover, this result shows that molecular hydrogen 306

concentration is dominated by the methane dissociation processes more than by all the 307

secondary reactions producing H2 and occurring in the plasma (neutral-neutral, ion-molecule 308

and ion dissociative recombination reactions). 309

On the contrary, H concentration increases linearly with initial methane concentrations 310

(Figure 5). No asymptote is observed for initial methane concentrations higher than 5% as for 311

molecular hydrogen. This suggests that atomic Hydrogen production is more independent 312

from methane direct dissociation than molecular Hydrogen. Atomic Hydrogen is probably 313

more linked with other reactions as for example dissociative recombination of positive ions 314

(Plessis et al., 2010). 315

316

3.2.2. Ammonia 317

Ammonia NH3 is of great interest concerning prebiotic chemistry, providing reactive 318

nitrogen-containing species to Titan’s atmosphere. It has been detected and quantified by the 319

INMS instrument onboard Cassini and are suggested to be chemically correlated (Yelle et al., 320

2010). 321

In our case, the ammonia production signature at m/z=17 is partially hidden by the 322

contribution of the water signal of the mass spectrometer. In order to extract this information 323

from the data, we precisely quantified the I17/I18 ratio given by the only water signature in our 324

20

QMS, using all the spectra recorded without plasma, so in absence of ammonia. Thirty eight 325

spectra have been recorded and a statistical fluctuation of I17/I18 can be deduced. A maximum 326

error bar of ±2σ is calculated and plotted for each experimental condition in Figure 8. The 327

I17/I18 measurement outside this statistical water range corresponds thus to a significant 328

difference compared with the water ratio. A larger value of I17/I18 is consequently significant 329

for the production of ammonia when the plasma is on and contributing to I17. 330

An increase of the ammonia production with the initial amount of methane can actually be 331

observed on Figure 8, and becomes significant for methane concentrations higher than 5%. 332

An ammonia production is not discarded for methane concentration below 5% but it is not 333

important enough to be distinguished from the experimental uncertainty of the I17/I18 water 334

ratio. 335

336

21

337

0 2 4 6 8 10

0,44

0,48

0,52

0,56

0,60

Plasma ON Plasma OFF

I 17 /

I 18

[CH4]0 (%) 338

Figure 8: Intensity ratios I17/I18 as a function of the initial methane concentration (red) when the plasma is 339

ON, (black) when the plasma is OFF. The uncertainties calculated corresponds to the 2σσσσ statistical 340

experimental scattering, calculated with 3 to 8 experiments when the plasma is ON for each methane 341

percentage, and with 30 spectra when the plasma is OFF (statistics on the whole set of experiments, 342

whatever the initial methane percentage). 343

344

Ammonia formation in CH4-N2 plasmas is an important issue which has not been definitely 345

solved yet (Horvath et al., 2011; Mutsukura, 2001; Uyama et Matsumoto, 1989). In any case 346

the key is the production of NH radicals, which react with molecular hydrogen to produce 347

ammonia through the reaction 348

NH + H2 � NH3 (4) 349

This reaction explains that ammonia production is promoted by a hydrogen-rich gas mixture. 350

The NH radical is suggested in the literature to be both produced by gas phase mechanism, 351

involving ion chemistry: 352

22

N2++CH4 � N2H

+ + CH3 (5) 353

N2H++e-

�NH+N (6) 354

and by radical chemistry: 355

N+ H�NH (7) 356

NH+ H2�NH3 (8) 357

Some studies suggests that the radical pathway can in some conditions be catalysed on the 358

surface of the reactor (Touvelle et al., 1987). However, in our experimental device, we can 359

assume that reactions are mainly in the plasma volume and that this wall effect can be 360

neglected. The ratio volume/reactor surface is actually larger in our case than in the set up of 361

Touvelle et al., 1987. 362

In addition to the identified and well-known mechanisms listed above, dissociative 363

recombination with electron of protonated amines may also provide plausible pathways in the 364

plasma towards ammonia, as the one suggested by Yelle et al., 2010 for Titan’s atmospheric 365

chemistry from protonated methanimine dissociative recombination (eq. 11). 366

We are in need of further experimental data on the dissociative recombination of nitrogeneous 367

ion species to evaluate this possible ammonia source in both Titan’s ionosphere and our 368

plasma experiment. 369

370

371

23

3.2.3. Organic species production 372

Figure 6 illustrates the rich product formation in the stationary plasma for an initial methane 373

concentration of 10%: a production of species can be observed in the ranges m/z=[24-29], m/z 374

= [36-44] and m/z = [50-55]. These mass ranges correspond respectively to elemental formula 375

of CxHyNz, x+z=2, CxHyNz, x+z=3,. CxHyNz, x+z=4. Those will be named C2, C3 and C4 376

blocks in the following. Ionization of a complex mixture of neutrals by electronic impacts 377

with electrons at energy of 70 eV leads to overlapping cracking patterns in the mass spectra, 378

which are commonly reported as difficult to interpret and set up a recurring issue for spectra 379

analysis. Only the main univocal mass peaks are often treated (Deschenaux et al., 1999; 380

Majumdar et al., 2005). 381

382

3.2.3.1. General evolution with methane initial amount 383

For this reason, we first analysed the evolution of the mass intensities according to the 384

methane amount, without any temptative identification. Results are plotted in a semi log scale 385

on Figure 9:, with 2σ uncertainties on the measurements calculated by a four times repetition 386

of each spectrum measurement. 387

388

24

0 2 4 6 8 101E-4

1E-3

0,01

0,1

Inte

nsity

(A

U)

[CH4]0 (%)

28

27

26

29

25

C2-block

0 2 4 6 8 10

1E-5

1E-4

1E-3

C3-block

38

44

37

Inte

nsity

(A

U)

[CH4]0 (%)

41

43

0 2 4 6 8 101E-5

1E-4

1E-3

C4-block

Inte

nsity

(A

U)

[CH4]0 (%)

5251

54

5053

Figure 9: Evolution of significant peaks detected, according to the initial methane amount. Reported 389

uncertainties are 2σ calculated on four repetitions of each spectrum and two experiments per methane 390

condition. 391

25

392

Globally, we observe an important effect of the initial amount of methane in the gas mixture, 393

with more and more gaseous species, also with larger intensities, detected when the initial 394

methane concentration increases. Less products peaks are observed in the mass spectrum for 395

the 1% methane condition, in comparison with the 10% methane one. This result is in 396

agreement with the previous study of Sciamma-O’Brien et al. (2010) on the gas-to-solid 397

conversion efficiency. It was shown that this efficiency decreases drastically with the initial 398

methane concentration, meaning that for similar methane consumption, the gas mixture 399

becomes more and more favourable to gaseous products in the detriment of solid products. 400

401

3.2.3.2. Identification of the species 402

Differential mass spectra for all the detected peaks (ISS(X)-I0(X)) have been measured for 403

[CH4]0 = 1 to 10%. Results obtained at 1, 5 and 10 % of methane in the initial gas mixture are 404

plotted in Figure 10. 405

406

26

407

20 25 30 35 40 45 50 55 60-6

-5

-4

-3

-2

-1

0

Log(

I SS- I 0)

m/z

1% CH4

5% CH4

10% CH4

408

Figure 10: Differential mass spectra (Stationary spectrum – Initial spectrum) for initial gas mixtures of 409

(black) 1%, (blue) 5% and (red) 10% of methane in the initial gas mixture. 410

411

Differential mass spectra for [CH4]0=1%: explicit treatment. 412

We observe that for this initial methane concentration, fewer molecules are produced in the 413

plasma. The identification is less ambiguous and the spectra can be treated in details. Mass 414

peaks can be singly attributed as follows. 415

The heaviest detected block, of C4 compounds is dominated by the peak at m/z = 52 416

and is attributed to C2N2, which has only one other fragment in the NIST database at m/z = 26 417

(5% of the m/z = 52 intensity). C4H4 is not possible because its fragments at 50 and 51 amu, 418

which are important in the NIST database (50% of the main m/z = 52 peak) are in 419

disagreement with the observed mass spectrum. C2N2 comes from the recombination of two 420

CN radicals. C2N2 has previously been observed in Titan's atmosphere (Cui et al., 2009; 421

27

Robertson et al., 2009; Vuitton et al., 2007) and in N2-CH4 plasma experiments (Bernard et 422

al., 2003; Majumdar et al., 2005; Ramírez et al., 2005). CN radicals have moreover been 423

detected by optical emission spectroscopy in N2-CH4 plasmas (Alcouffe et al., 2010; Horvath 424

et al., 2011). 425

The C3 block includes 4 main peaks at mass m/z 44, 43, 41 and 38. 426

Propane (C3H8) is the most probable product at m/z 44. The odd masses at m/z = 41 and 43 427

can be attributed to one nitrogen-bearing species. Acetonitrile (CH3CN) is unambiguously at 428

m/z = 41, and has been cross-confirmed as a major product in the experiment by the GCMS 429

analysis of Gautier et al., 2011. The mass peak at m/z = 43 is the signature of ethanimine 430

CH3-CH=NH (and/or amino-ethene with the unsaturation carried by the C=C bond: CH2=CH-431

NH2), which is cross-confirmed by the ion detection of Mutsukura, 2001 (see section 3.3). 432

Ethanimine has been shown to be produced by the reaction between N(2D) and C2H4 433

(Balucani et al., 2010). However, it is not to exclude that this mass peak may also be the result 434

of a fragmentation pattern of amino-methanimine (H2N-CH=NH), its fragmentation pattern 435

being unknown. However, the ion study of Mutsukura, 2001 does not show the signature of 436

amino-methanimine. Further studies with complementary in situ measurements could be 437

necessary to unveil the ambiguity. 438

The m/z 38 mass peak is observed in all the reactive N2-CH4 studies made by mass 439

spectrometry (Imanaka et Smith, 2010; Majumdar et al., 2005), but there is not clear evidence 440

to attribute this fragment yet. A C2N+ fragment from the C2N2 neutral can be suspected. 441

442

The C2 block shows significant peak intensities for m/z = 25, 26, 27, 28 and 29. They are 443

in agreement with the similar N2-CH4 plasma studies performed by Majumdar et al., 2005. 444

Three main points can be extracted from these C2 block data, largely complicated by the 445

overlapping of several hydrocarbon and nitrogenous containing fragments. 446

28

• The HCN formation (main peak at m/z = 27), is cross identified as the major 447

gaseous products by Gautier et al., 2011 in the trap analysis. 448

• An important intensity of the m/z = 28 peak, corresponds mainly to ethylene C2H4. 449

Indeed, as we analyse the differential spectrum between final and initial states, the 450

increase of the m/z 28 peak cannot be due to a production of N2. 451

• The production of acetylene C2H2 at m/z = 26 is possible but its main fragment at 452

m/z = 26 is mixed with the contributions of the CN+ fragments of all abundant 453

nitriles. 454

A firm detection of methanimine, CH2=NH at m/z = 29 can be given for [CH4]0=1%. Actually 455

an additional contribution at m/z = 29 can be given by the propane fragmentation 44(50) 456

29(100) pattern (database of the Pfeiffer QMS 200). The ratio of these two peaks in the 457

differential spectra is therefore plotted Figure 11 as a function of initial percentage of 458

methane. For [CH4]0 larger than 2%, the ratio is constant and consistent with the propane 459

pattern, showing that this compound becomes the main contributor to the m/z = 29 peak for 460

these methane amounts. On the other hand, the ratio is much larger for [CH4]0=1%, and 461

becomes inconsistent with propane. It confirms the important contribution of methanimine at 462

m/z = 29 in these experimental conditions. Methanimine was previously detected in a N2-CH4 463

microwave plasma discharge by Fujii et Arai, 1999; Fujii et Kareev, 2001; Kareev et al., 464

2000. However, micro-wave discharges are well-known to be very warm and to be compared 465

with combustion processes. Our study confirms thus the methanimine production in much 466

cooler experimental conditions, more representative of the cold Titan’s atmospheric 467

chemistry. Methanimine has been detected and quantified by the INMS instrument onboard 468

Cassini (Yelle et al., 2010) and is mainly produced by the reaction between N(2D) and 469

methane (Balucani et al., 2009). 470

29

0 2 4 6 8 100

2

4

6

8

10

[Iss

(29)

- I 0(2

9)] /

[Iss

(44)

- I 0(4

4)

[CH4]0, %

471

Figure 11: Evolution of the ratio between the increases of m/z 44 and m/z 29 when the plasma is ON. Two 472

experiments were performed per [CH4]0 conditions. The stable value of about 2 for [CH4]0 concentration 473

larger than 2% is in agreement with a co-evolution of fragments intensities for propane. However, the 474

larger value of this ratio for [CH 4]0=1% indicates an additional contribution of methanimine at mass 29 u. 475

476

Table 1 summarizes the identified molecules from in the differential mass spectra at 477

[CH4]0=1%. 478

C2 block

m/z 26 27 28 29

Molecule

HCN

(CN

fragment)

HCN C2H4 CH2=NH

C3 block m/z 38 41 43 44

Molecule C2N2 ? CH3CN CH3-CH=NH C3H8

C4 block m/z 52 - - -

Molecule C2N2 - - -

Table 1: Molecules identified in the differential mass spectra at [CH4]0=1%. 479

30

Gas mixture with [CH4]0 > 1 %: Single Value Decomposition. 480

The spectra obtained at higher amount of methane in nitrogen are more complex than in the 481

1% case. Single-Value Decomposition (SVD) algorithms have been used for the treatment of 482

the Cassini INMS neutrals mass spectra (Cui et al., 2009) and of a photochemical 483

experimental simulation of Titan's atmosphere (Imanaka et Smith, 2007). These systematic 484

methods improve notably the processing of the spectra, but their non unique decomposition 485

results are also known to be very dependent on the chosen database and to be taken with a lot 486

of care. Studies on quantification and identification were performed by Robert, 2010; Turner 487

et al., 2004. They showed that QMS instruments with electron impact ionization methods 488

have strong non-linearities, and cross-sensitivity effects (dependence of the response of a 489

compound according to its co-mixture with other species). A proper calibration is almost 490

impossible for gas mixture more complex than 5 species. In our case, the GC-MS study of 491

Gautier et al., 2011, has revealed a gas mixture of about 40 different species, often non-492

commercially available. We therefore made use of such decomposition tools in order to go 493

further in the understanding of the gas mixture composition, but we have been very careful 494

not to over-interpret the data inversion. 495

Our reference database was built according to the following rules. We used preferentially the 496

specific Pfeiffer database provided with the instrument, which is significantly different from 497

the NIST (Stein, 2011) fragmentation patterns for ionization by a 70 eV electronic impact. We 498

found for N2 and CH4 some differences between the NIST and the Pfeiffer database of about 499

20-30% on the fragmentation patterns. For species absent from the Pfeiffer database but 500

present in the NIST, the reference mass spectra were taken from the NIST. 501

The SVD decomposition was limited on the mass range above m/z 20, for CxHyNz species 502

with x+z ≥ 2, adapted to products identification. As in Imanaka et Smith, 2010, species were 503

added progressively and a minima. 504

31

A few clear conclusions could however be obtained. 505

• The main peaks of the spectrum can be easily fitted with nitriles: HCN (27 u), CH3CN 506

(41 u), C2H3CN (53 u), C2H5CN (55 u), HC3N (51 u) and C2N2 (52 u) and are 507

consistent with the gas-phase studies of Coll et al., 1999; De Vanssay E et Raulin, 508

1995; Gautier et al., 2011; Ramírez et al., 2001; Ramírez et al., 2005. 509

• Hydrocarbons have complex fragmentation patterns and do not contribute explicitly to 510

specific and univocal mass peaks. We were therefore only able to confirm that their 511

introduction is useful to improve the fit, but they are not explicitly constrained by the 512

mass spectrum and cannot be strictly identified. As a consequence, we suspect that the 513

attribution made in the comparable study of Imanaka et Smith, 2010, provides also 514

estimations but cannot be considered as a firm detection. 515

• Moreover, no hydrocarbon allows to correctly simulate the peaks m/z = 54, 38, 37 and 516

36. Because of the fragmentation pattern description weaknesses, we cannot conclude 517

that a species is missing to explain masses 36, 37 and 38 u. Indeed they are located on 518

the lower masses of the C3 block. They also could be due to the fragments of larger 519

hydrocarbons, poorly taken into account. However, it is not the case for the mass 54 520

which is located on the right side of the C4 block. The impossibility to fit the mass 521

spectrum with well-known hydrocarbons suggests a nitrogen bearing species, not yet 522

taken into account in the database. The condensation of HCN and methanimine 523

(H2C=N-CN or HN=CH-CN) could possibly explain such an m/z 54 peak. 524

525

3.2.3.3. Role of hydrocarbons and nitrogen-containing species in the organic growth. 526

It is more or less admitted that Titan’s atmospheric aerosols may possess a polymeric 527

structure (at least for the aerosol nucleation step), but their structural basis remains unknown 528

because no instrument onboard Cassini provides a precise chemical information on the 529

32

aerosol composition. The only ACP instrument (Israel et al., 2005) of the Huygens probe 530

provides qualitative analysis of the global chemical content of the aerosols in Titan’s 531

atmosphere. It used pyrolysis to analyse the refractory nucleus of aerosols during its descent 532

through Titan’s atmosphere. The major pyrolysis products were HCN and NH3 showing that 533

Titan’s aerosols are in fact very nitrogen-rich, which excludes the commonly suggested 534

theory that these aerosols are derived from polycyclic aromatic hydrocarbons (PAHs). 535

PAMPRE tholins polymeric structure has been studied in Carrasco et al., 2009; Pernot et al., 536

2010, who confirm an important contribution of nitrogen in the aerosols composition as in 537

Titan’s aerosols; and its central role in the unsaturated component of tholins. 538

Nevertheless, a few studies on Titan’s atmosphere (Imanaka et Smith, 2007; Waite Jr. et al., 539

2007) and on pure methane and CH4-N2 plasmas (Deschenaux et al., 1999; Majumdar et al., 540

2005) emphasize a chemical growth carried by hydrocarbons (through ionic or neutral 541

processes) which could even been scavenged in the presence of nitrogen (Majumdar et al., 542

2005). 543

Those two hydrocarbon and nitrogenous pathways are not exclusive and possibly both 544

participate to the chemical blocks constitutive of Titan’s aerosols. Nevertheless, it would be 545

hazardous to reduce organic aerosol formation in the solar system to the only PAH theory just 546

because the nitrogen chemistry is far from being understood (see for example the recent paper 547

of Teanby et al., 2010 highlighting an unsuspected chemistry of nitriles). It is important to 548

remind that Quirico et al., 2008 found in an IR and Raman study of the PAMPRE tholins a 549

signature of nitrogen-rich aromatic rings: triazine (C3N3 aromatic ring), but none of benzene 550

aromatic ring. This was also confirmed by Gautier et al. (2011) where the analysis of the 551

PAMRE gas phase by GC-MS revealed that most abundant aromatics compounds are nitrogen 552

bearing ones, and that pure hydrocarbons aromatics are only minor compounds. Efforts on 553

33

this original N-PAH chemistry should be given in the future and could impact larger scales 554

than the specific Titan’s aerosols. 555

Our study confirms the importance of N-bearing compounds for the production of solid 556

particles simulating a Titan-like partially ionized N2-CH4 gas mixture: the PAMPRE reactive 557

plasma, ionized at the order of 10-2 ppm (Alcouffe et al., 2010), less than in Titan’s 558

ionosphere (0.1-1 ppm), is full of nitrogen-rich solid particles and exhibits a gas mixture 559

largely dominated by N-bearing gas products, including nitriles, ammonia and imines. 560

561

562

34

3.3. Positive ion species 563

Ammonia NH3 and methanimine CH2=NH are of great interest concerning prebiotic 564

chemistry, providing reactive nitrogen-containing species to Titan’s atmosphere. 565

In the ionosphere, the protonation, of methanimine followed by dissociative recombination of 566

the ion with electrons (reaction 11) may explain an additional production of NH2, and all the 567

more of ammonia in Titan ionosphere according to the following mechanism proposed by 568

Yelle et al. (2010). 569

N(2D)+CH4�CH2=NH+H (9) 570

CH2=NH+XH+�CH2=NH2

++X (10) 571

CH2=NH2++e-

�CH2+NH2 (11) 572

NH2+H2CN�NH3+HCN (12) 573

Methanimine is detected by mass analysis of the neutral species in our experiment for 574

[CH4]0=1% (see Section 3.2), but its signature at m/z=29 is hidden by a propane fragment for 575

the other methane conditions (see Figure 11). 576

In order to have a confirmation of this methanimine production at the larger concentration of 577

methane, we analyzed the positive ion study performed by Mutsukura, 2001 in a similar RF 578

plasma setup but working at lower pressure (0.13 mbar) than in our case. Methanimine 579

production can actually be inferred from the measurement of its protonated form CH2NH2+ 580

(unambiguous signature at m/z 30), as it has been done for Titan’s ionosphere with the ion 581

INMS Cassini spectra (Yelle et al., 2010). Mutsukura measured the ions characteristic of a 582

pure nitrogen, a pure methane and a 50% nitrogen-methane radio-frequency plasma. The mass 583

spectra are given in arbitrary units on Figure 12a. 584

585

586

35

10 15 20 25 30 35 40 45 50 55 600

102030405060708090

100

Inte

nsity

, a.u

m/z

CH5

+

10 15 20 25 30 35 40 45 50 55 600

200400600800

10001200140016001800

CH3CHNH

2

+

CH3CNH+

CH2NH

2

+

HCNH+

NH4

+

(b)

RF Plasma N2/CH

4 (50/50)

Plasma N2

Plasma CH4

Plasma N2/CH

4 (50/50)

INMS T19

Inte

nsity

, a.u

m/z

(a)

587

Figure 12: (a) Ion mass spectra measured in an RF plasma in pure nitrogen, pure methane and a 50-50 588

gas mixture and adapted from the study of Mutsukura, 2001. (b) Qualitative comparison of positive ion 589

mass spectra observed in a N2-CH4 RF plasma and in Titan’s ionosphere, in arbitrary units and 590

normalized at m/z 28. 591

592

36

Figure 12-b is a comparison of the ion mass spectra measured both in Titan’s ionosphere 593

(Waite Jr. et al., 2007) and in the RF-N2-CH4 plasma of Mutsukura, 2001. It shows a 594

qualitative agreement between the plasma setup and Titan’s ionosphere: the ions detected in 595

both ionized media are the same. However it also illustrates a quantitative difference: in the 596

Mutsukura’s RF plasma generates much more primary ions in comparison with heavier ions 597

than Titan’s ionosphere. Titan’s ionosphere on the contrary exhibits a larger proportion of 598

heavy ions. The difference can be explained by a more efficient production of primary ions in 599

the RF plasma, because the electron energy distribution function in the plasma has a longer 600

tail in the higher energies responsible for the primary ions production than the solar spectrum 601

(Szopa et al., 2006). In laboratory plasmas, positive ions and electrons are lost by 602

recombination and diffusion to the metallic walls. The primary ions are therefore less 603

available for ion-molecule reactions, processes responsible for the ion growth in Titan’s 604

ionosphere as shown in Carrasco et al., 2008. 605

As in the INMS ion mass spectra, an intense peak at m/z 30 is highlighted in the RF nitrogen-606

methane plasma which is absent in the pure and in the methane plasmas. Methanimine is 607

therefore a characteristic product of RF N2-CH4 plasmas. 608

609

3.3.1. Discussion on the methanimine production pathways 610

The simultaneous production of NH3 and methanimine is in agreement with the correlation 611

proposed by Yelle et al.(2010) for Titan’s ionosphere, but it is not clear whether ammonia is a 612

source for methanimine, or the opposite, as suggested by Yelle et al. Indeed different sources 613

can be found in the literature concerning the possible production pathways of these two 614

nitrogenous volatiles. Those are detailed below. 615

We identified two types of production pathways for methanimine compatible with Titan’s 616

ionosphere and with PAMPRE plasma chemistry: 617

• A radical pathway (Balucani et al., 2009) 618

N(2D)+CH4 � CH2=NH+H (13) 619

37

In this case, the positive correlation between methanimine production and methane 620

concentration is trivial. 621

• Ion chemistry pathways forming protonated methanimine, CH4N+, mainly involving 622

hydrocarbon ions with ammonia (Carrasco et al., 2008, based on the reference 623

database of Anicich, 2003) 624

o CH2+ + NH3 � CH4N

+ + H (14) 625

o CH3+ + NH3 � CH4N

+ + H2 (15) 626

o C3H6+ + NH3 � CH4N

+ + C2H5 (16) 627

o NH2+ + C2H4 � CH4N

+ + 1CH2 (17) 628

The rate constants of the above ion-molecule reactions are all in the same order of magnitude 629

of about 10-9 cm3.s-1 (Langevin rate constant for ion/molecule capture). However we can 630

suspect an important role of the reaction involving methyl ion CH3+ with ammonia, because 631

of the efficient production pathways of CH3+. This ion is indeed major in the nitrogen-632

methane plasma measured by Mutsukura, 2001 (Figure 12: a), and its production involves the 633

reactions of the two main nitrogen ions N2+ and N+ with methane. Moreover we can note that 634

CH3+ production is favoured by an increase of the methane concentration, which would 635

suggest an increase of the observed methanimine production with methane. The co-reactant 636

with CH3+ leading to methanimine is ammonia, which is also positively correlated with the 637

methane concentration through the N++CH4 reaction. 638

Methanimine ionic production pathways depend on the recombination of protonated 639

methanimine with electrons. There is however no data on the fragmentation pattern of this ion 640

by dissociative recombination with electrons (Plessis et al., 2010), to confirm that 641

methanimine could actually be the major product of such a recombination. It is precisely for 642

this reaction that Yelle et al., 2010 estimated two balanced pathways to explain ammonia 643

38

production in Titan’s ionosphere: a 0.5 H-loss branching ratio leading to methanimine 644

production, and a 0.5 dissociative branching ratio leading to NH2 production. 645

646

3.3.2. Impact of the detection of methanimine on Titan’s aerosol production 647

Methanimine is known as an easily polymerizable compound and is already suspected to take 648

part in Titan’s aerosol formation (Balucani et al., 2010). An oligomer of methanimine would 649

actually be in agreement with the C/N alternance detected in the paper of Carrasco et al., 2009 650

where a C2N3- common systematic feature was identified in the fragmentation pattern of all 651

the negative ions obtained after APPI ionization of tholins. It is also in agreement with the 652

result that all the double bonds in tholins are supported by C=N bonds, as shown in Pernot et 653

al., 2010; Somogyi et al., 2005. 654

Models describing the aerosols production and evolution in Titan’s atmosphere involve 655

polyynes, aromatics and nitrile pathways (Lavvas et al., 2008; Lebonnois et al., 2002). It 656

would be important for future work on Titan’s aerosols micro-physics to also implement this 657

newly discovered imine pathway. 658

659

39

4-Conclusion 660

The study of the stationary gas phase composition obtained with various initial methane 661

concentrations from 1 to 10%, confirms the enrichment of the gas phase - number of species 662

detected and global quantity - with the input of methane observed in Gautier et al., 2011. A 663

specific case is found at 1%, which almost only produces nitrogenous species and few 664

hydrocarbons. Atomic hydrogen is quantified and an anti-correlation is found between the 665

abundance of atomic hydrogen and the aerosol production efficiency. This result is in 666

agreement with the inhibiting heterogeneous effect suggested in Sciamma-O'Brien et al., 667

2010. A modelling of nitrogen-methane RF plasma, similar to the study of Pintassilgo et al., 668

1999 but which was dedicated to a N2-CH4 glow discharge, would be an essential 669

complementary tool enabling to interpret the correlations found here between volatiles and 670

aerosol production. 671

Moreover ammonia and methanimine are detected in the plasma in agreement with the 672

detection by INMS in Titan’s ionosphere (Yelle et al., 2010), such as ethanimine. The 673

production in the gas phase of such highly polymerizable species provides promising clues to 674

understand the production of tholins. 675

In the frame of Titan’s upper atmosphere, our study confirms two main antagonist effects on 676

aerosol production and growth. The first effect supports the organic growth process suggested 677

by Balucani et al. (2009, 2010) and involving a co-polymerization with imine species. The 678

second one is an inhibiting effect of atomic hydrogen on the aerosol production. This effect is 679

related to the observation of Sekine et al. (2008a, 2008b), showing an efficient fixation of 680

atomic hydrogen on aerosols. Here we show that the efficient hydrogen fixation is made at the 681

expense of the aerosol organic growth itself and has to be considered as a competition with 682

the aerosol organic growth process. 683

40

Acknowledgements 684

We thank L. Hogrel for her participation to the experimental work. We acknowledge the 685

CNRS (PID OPV, PNP), and the PRES UniverSud and the French National Research Agency 686

(ANR-09-JCJC-0038) for their financial support. 687

688

41

Figure 1: Blank mass spectrum of the mass spectrometer only. ................................................ 7 689

Figure 2: Comparison between a scan analog and a scan bargraph acquisition on the example 690

of the stationary state obtained after plasma discharge in a 1% methane initial gas mixture. ... 8 691

Figure 3: Typical example of emission spectrum obtained between 475 and 828 nm for 5% 692

CH4/N2 RF plasma with the presence of 5% Ar as an actinometer gas. The mean features 693

emissions are identified together their wavelengths. ............................................................... 11 694

Figure 4: Dependence of (a) Hβ (486.1 nm) and (b) Ar (811.5 nm) emission line intensities on 695

the percentage of CH4 added to N2 RF plasma. The percentage of Ar as an actinometer gas is 696

5% for all measurements. ......................................................................................................... 13 697

Figure 5: Evolution of hydrogen atom densities as a function of % CH4 obtained from Figure 698

4 in the same experimental conditions. .................................................................................... 14 699

Figure 6: Mass spectrum of a CH4-N2 10-90% gas mixture, (black) plasma OFF, (red) plasma 700

ON. ........................................................................................................................................... 16 701

Figure 7: Hydrogen production (difference between the stationary Iss(2) and the initial I0(2) 702

intensities of the mass peak at m/z=2) (above) according to the initial methane concentration 703

in the gas mixture: (below) according to the methane consumption (difference between the 704

stationary Iss(15) and the initial I0(15) intensities of the mass peak at m/z=15). ..................... 18 705

Figure 8: Intensity ratios I17/I18 as a function of the initial methane concentration (red) when 706

the plasma is ON, (black) when the plasma is OFF. The uncertainties calculated corresponds 707

to the 2σ statistical experimental scattering, calculated with 3-8 experiments when the plasma 708

is ON for each methane percentage, and with 30 spectra when the plasma is OFF (statistics on 709

the whole set of experiments, whatever the initial methane percentage). ................................ 21 710

Figure 9: Evolution of significant peaks detected, according to the initial methane amount. 711

Reported uncertainties are 2σ calculated on four repetitions of each spectrum and two 712

experiments per methane condition. ......................................................................................... 24 713

42

Figure 9: Differential mass spectra (Stationary spectrum – Initial spectrum) for initial gas 714

mixtures of (black) 1%, (blue) 5% and (red) 10% of methane in the initial gas mixture. ....... 26 715

Figure 11: Evolution of the ratio between the increases of m/z 44 and m/z 29 when the plasma 716

is ON. The stable value of about 2 for [CH4]0 concentration larger than 2% is in agreement 717

with a co-evolution of fragments intensities for propane. However, the larger value of this 718

ratio for [CH4]0=1% indicates an additional contribution of methanimine at mass 29 u. ........ 29 719

Figure 12: (a) Ion mass spectra measured in an RF plasma in pure nitrogen, pure methane and 720

a 50-50 gas mixture and adapted from the study of Mutsukura, 2001. (b) Qualitative 721

comparison of positive ion mass spectra observed in a N2-CH4 RF plasma and in Titan’s 722

ionosphere, in arbitrary units and normalized at m/z 28. ......................................................... 35 723

724

725

726

727

728

729

730

43

References 731

Alcouffe, G., et al., 2010. Capacitively coupled plasma used to simulate Titan's atmospheric chemistry. 732 Plasma Sources Science and Technology 19, 015008. 733

Anicich, V., 2003. An Index of the Literature for Bimolecular Gas Phase Cation-Molecule Reaction 734 Kinetics. JPL Publication 03-19, 1-1194. 735

Balucani, N., et al., 2009. Combined Crossed Molecular Beam and Theoretical Studies of the N(2D) + CH4 736 Reaction and Implications for Atmospheric Models of Titan. The Journal of Physical Chemistry 737 A. 113, 11138-11152. 738

Balucani, N., et al., 2010. Formation of nitriles and imines in the atmosphere of Titan: combined crossed-739 beam and theoretical studies on the reaction dynamics of excited nitrogen atoms N(2D) with 740 ethane. Faraday Discussions. 147, 189-216. 741

Bernard, J.-M., et al., 2003. Experimental simulation of Titan’s atmosphere: Detection of ammonia and 742 ethylene oxide. . Planet. Space Sci. 51, 1003-1005. 743

Carrasco, N., et al., 2008. Toward a reduction of the bimolecular reaction model for titan's ionosphere. 744 International Journal of Chemical Kinetics. 40, 699-709. 745

Carrasco, N., et al., 2009. Chemical Characterization of Titan's Tholins: Solubility, Morphology and 746 Molecular Structure Revisited. . The Journal of Physical Chemistry A. 113, 11195-11203. 747

Coll, P., et al., 1999. Experimental laboratory simulation of Titan’s atmosphere: Aerosols and gas phase. 748 Planet. Space Sci. 47, 1331-1340. 749

Cui, J., et al., 2009. Analysis of Titan's neutral upper atmosphere from Cassini Ion Neutral Mass 750 Spectrometer measurements. Icarus. 200, 581-615. 751

De Vanssay E, G. M. C. G. J. C., Raulin, F., 1995. Experimental simulation of Titan's organic chemistry at 752 low temperature. Planet. Space Sci. 43, 25. 753

Deschenaux, C., et al., 1999. Investigations of CH4 , C2H2 and C2H4 dusty RF plasmas by means of FTIR 754 absorption spectroscopy and mass spectrometry. Journal of Physics D: Applied Physics. 32, 1876. 755

Fujii, T., Arai, N., 1999. Analysis of N-containing hydrocarbon species produced by a CH4/N2 microwave 756 discharge: simulation of Titan's atmosphere. The Astrophysical Journal. 519, 858-863. 757

Fujii, T., Kareev, M., 2001. Diagnostic studies of a CH4/N2 microwave plasma by mass spectrometry: ionic 758 and neutral species. J. Phys. Chem. A. 105, 4923-4927. 759

Gautier, T., et al., 2011. Nitrile gas chemistry in Titan atmosphere. Icarus. 213, 625-635. 760 Gicquel, A., et al., 1998. Validation of actinometry for estimating relative hydrogen atom densities and 761

electron energy evolution in plasma assisted diamond deposition reactors. J. Applied Physics. 83, 762 7504-21. 763

Horvath, G., et al., 2011. Organic chemistry of NH3 and HCN induced by an atmospheric abnormal glow 764 discharge in N2-CH4 mixtures. The European Physical Journal - Applied Physics. 53. 765

Imanaka, H., Smith, M. A., 2007. Role of photoionization in the formation of complex organic molecules in 766 Titan's upper atmosphere. Geophys. Res. Lett. 34, L02204. 767

Imanaka, H., Smith, M. A., 2010. Formation of nitrogenated organic aerosols in the Titan upper 768 atmosphere. Proceedings of the National Academy of Sciences. 107, 12423-12428. 769

Israel, G., et al., 2005. Complex organic matter in Titan’s atmospheric aerosols from in situ pyrolysis and 770 analysis. Nature. 438. 771

Kareev, M., et al., 2000. Diagnosis of a CH4/N2 microwave discharge: ionic and neutral species. J. Phys. 772 Chem. A. 104, 7218-7223. 773

Lavvas, P. P., et al., 2008. Coupling photochemistry with haze formation in Titan's atmosphere, Part II: 774 Results and validation with Cassini/Huygens data. Planetary and Space Science. 56, 67-99. 775

Lebonnois, S., 2005. Benzene and aerosol production in Titan and Jupiter's atmospheres: a sensitivity 776 study. Planetary and Space Science. 53, 486-497. 777

Lebonnois, S., et al., 2002. Transition from Gaseous Compounds to Aerosols in Titan's Atmosphere. 778 Icarus. 159, 505-517. 779

Majumdar, A., et al., 2005. Chemical Reaction Studies in CH4/Ar and CH 4/N2 Gas Mixtures of a Dielectric 780 Barrier Discharge. The Journal of Physical Chemistry A. 109, 9371-9377. 781

Mutsukura, N., 2001. Deposition of Diamondlike Carbon Film and Mass Spectrometry Measurement in 782 CH4/N2 RF Plasma. Plasma Chemistry and Plasma Processing. 21, 265-277. 783

Pernot, P., et al., 2010. Tholinomics: Chemical Analysis of Nitrogen-Rich Polymers. Analytical Chemistry. 784 82, 1371-1380. 785

Pintassilgo, C. D., Loureiro, J., 2010. Kinetic study of a N2-CH4 afterglow plasma for production of N-786 containing hydrocarbon species of Titan's atmosphere. Advances in Space Research. 46, 657-671. 787

44

Pintassilgo, C. D., et al., 1999. Methane decomposition and active nitrogen in a N2 -CH4 glow discharge at 788 low pressures. Plasma Sources Science and Technology. 8, 463. 789

Plessis, S., et al., 2010. Knowledge-based probabilistic representations of branching ratios in chemical 790 networks: The case of dissociative recombinations. The Journal of Chemical Physics. 133, 134110-791 21. 792

Quirico, E., et al., 2008. New experimental constraints on the composition and structure of tholins. Icarus. 793 198, 218-231. 794

Ramírez, S. I., et al., 2001. Possible contribution of different energy sources to the production of organics 795 in Titan's atmosphere. Advances in Space Research. 27, 261-270. 796

Ramírez, S. I., et al., 2005. Organic chemistry induced by corona discharges in Titan's troposphere: 797 Laboratory simulations. Adv. Space Res. . 36, 274-280. 798

Robert, E., 2010. Mass spectrometer calibration over wide concentration ranges in multicomponent gas 799 mixtures. Meas. Sci. Technol. 21, 1-11. 800

Robertson, I. P., et al., 2009. Structure of Titan's ionosphere: Model comparisons with Cassini data. 801 Planetary and Space Science. 57, 1834-1846. 802

Sciamma-O'Brien, E., et al., 2010. Titan's atmosphere: An optimal gas mixture for aerosol production? 803 Icarus. 209, 704-714. 804

Sekine, Y. et al., 2008a. The role of organic haze in Titan’s atmospheric chemistry I. Laboratory 805 investigation on heterogeneous reaction of atomic hydrogen with Titan tholin. Icarus 194, 186–806 200. 807

Sekine, Y. et al., 2008b. The role of organic haze in Titan’s atmospheric chemistry II. Effect of 808 heterogeneous reaction to the hydrogen budget and chemical composition of the atmosphere. 809 Icarus 194, 201–211. 810

Somogyi, A., et al., 2005. Organic Environments on Saturn’s Moon, Titan: Simulating Chemical Reactions 811 and Analyzing Products by FT-ICR and Ion-Trap Mass Spectrometry. American Society for 812 Mass Spectrometry. 16, 850-859 813

Stein, S. E., "Mass Spectra" in the NIST Chemistry WebBook 69. http://webbook.nist.gov. In: P. J. 814 Linstrom, W. G. Mallard, Eds.), National Institute of Standards and Technology, Gaithersburg 815 MD, 20899, 2011. 816

Szopa, C., et al., 2006. PAMPRE: A dusty plasma experiment for Titan’s tholins production and study. 817 Planet. Space Sci. 54, 394-404. 818

Teanby, N. A., et al., 2010. Mapping Titan's HCN in the far infra-red: implications for photochemistry . 819 Faraday Discussions. 147, 51-64. 820

Touvelle, M., et al., 1987. Plasma Chemical Synthesis. II. Effect of Wall Surface on the Synthesis pf 821 Ammonia. Plasma Chemistry and Plasma Processing. 7, 101-108. 822

Turner, P., et al., 2004. Calibration effects during natural gas analysis using a quadrupole mass 823 spectrometer. Trends in Analytical Chemistry. 23, 281-287. 824

Uyama, H., Matsumoto, O., 1989. Synthesis of ammonia in high-frequency discharges. II. Synthesis of 825 ammonia in a microwave discharge under various conditions. Plasma Chemistry and Plasma 826 Processing. 9, 421-432. 827

Vuitton, V., et al., 2007. Ion chemistry and N-containing molecules in Titan's upper atmosphere. Icarus. 828 191, 722-742. 829

Waite Jr., J. H., et al., 2007. The Process of Tholin Formation in Titan ’s Upper Atmosphere. Science. 316, 830 870-875. 831

Yelle, R. V., et al., 2010. Formation of NH3 and CH2NH in Titan's upper atmosphere. Faraday 832 Discussions. 147, 31-49. 833

834 835