Sei sulla pagina 1di 112

 

 
Plasma diagnostics for the low-pressure plasma polymerization process: A
critical review

Damien Thiry, Stephanos Konstantinidis, Jérôme Cornil, Rony Snyders

PII: S0040-6090(16)00157-7
DOI: doi: 10.1016/j.tsf.2016.02.058
Reference: TSF 35058

To appear in: Thin Solid Films

Please cite this article as: Damien Thiry, Stephanos Konstantinidis, Jérôme Cornil,
Rony Snyders, Plasma diagnostics for the low-pressure plasma polymerization process:
A critical review, Thin Solid Films (2016), doi: 10.1016/j.tsf.2016.02.058

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Plasma diagnostics for the low-pressure plasma


polymerization process: a critical review

PT
Damien Thiry1,2, Stephanos Konstantinidis1, Jérôme Cornil3 and Rony Snyders1, 4

RI
1) Chimie des Interactions Plasma-Surface, CIRMAP, University of Mons, Place du Parc 20,

SC
B-7000 Mons, Belgium.

2) Institut des Matériaux Jean Rouxel, Université de Nantes, CNRS, 2 rue de la Houssinière

NU
B.P. 32229, 44322 Nantes cedex 3, France.

3) Service de Chimie des Matériaux Nouveaux (CMN), CIRMAP, University of Mons, 20


MA
Place du Parc, B-7000 Mons, Belgium.

4) Materia Nova Research Center, Parc Initialis, Avenue Nicolas Copernic 1, B-7000 Mons,
Belgium.
D
P TE
CE
AC

1
ACCEPTED MANUSCRIPT

Abstract

Since the 1980s, functionalized plasma polymer films have attracted a considerable

attention owing to their promising utilization in a wide range of modern applications. For such

PT
materials, controlling the chemistry of the coatings by a clever choice of the process parameters

RI
represents the main challenge. And yet, it became quickly obvious that in view of the complexity

SC
of the growth mechanism, fine control of the layers properties can only be reached by

understanding at a fundamental level the mechanistic formation of the layers. In this context, a

NU
detailed comprehensive study of plasma chemistry is therefore of crucial importance as the
MA
numerous interlinked chemical reactions occurring in the discharge govern the film properties. In

this paper, the most common plasma diagnostics methods employed in the context of plasma
D

polymerization process, namely Mass Spectrometry, in-situ Fourier Transform Infrared


TE

Spectroscopy, Optical Emission Spectroscopy, Langmuir and Ionic probes are reviewed. After a

light description of each technique, the main achievements for improving the mechanistic
P
CE

understanding of the layer formation are exposed. Moreover, the use of theoretical calculations

based on Density Functional Theory (DFT) to support the understanding of the acquired data is
AC

highlighted. In view of the better control of the process allowed by the plasma phase

investigation, some general conclusions and perspectives describing future developments in the

field of plasma polymerization are finally discussed.

Keywords: plasma polymerization, plasma polymer films, plasma diagnostic, growth mechanism

2
ACCEPTED MANUSCRIPT

1. Introduction

The interactions of a solid with its surrounding are mainly defined by the physico-chemical

properties of its surface. It is therefore not surprising that since many decades, continuous

PT
research and developments in materials science have contributed to the rapid growth of surface

RI
and coating technologies which nowadays still increasingly attract considerable attention. Surface

SC
technologies refer to the modification of the surface (e.g., chemical functionalization, etching,

coating,…) of a material without changing its bulk properties. For instance, tailor-made coatings

NU
allow adjusting mechanical (wear, friction), chemical (corrosion, permeation, temperature
MA
insulation, biocompatibility, wettability), electrical (conductivity), and optical (transmission,

reflection, absorption, color) properties of materials [1].


D

Through the years, numerous processes have been developed for the modification of surfaces
TE

via the synthesis of thin films. A non-exhaustive list includes chemical vapor deposition, pulsed
P

laser deposition, spin coating, sol gel, spin casting, thermal evaporation and plasma-based
CE

technologies [1-4]. Among them, the plasma-based processes are of particular interest by
AC

combining significant advantages such as their low process temperature, enabling the treatment

of a wide range of materials including polymers, and the absence of solvents making these

techniques compatible with the modern quest for environmentally friendly technology. Another

key advantage of these processes is their versatility enabling one to modulate the properties of a

given surface over a wide range (e.g., crystallinity, morphology, chemical composition of the

deposited material) by adjusting the synthesis conditions [5-9]. All these attractive properties

justify the popularity gained by plasma technologies and their important development in

numerous industrial fields such as automotive, aeronautics and microelectronics [1, 10, 11].

3
ACCEPTED MANUSCRIPT

If in the past, research and applications have often focused on the development of inorganic

thin films, the design of organic surfaces is nowadays more and more important with applications

in the fabrication of antibacterial surfaces [12], protein biochips [13, 14] or platforms for

PT
biomolecules immobilization [15, 16]. These surfaces can also be synthesized using plasma-

RI
based technology, more specifically by means of the plasma polymerization method, allowing the

formation of solid organic thin films referred to as plasma polymer films (PPF). Despite the use

SC
of the word “polymer”, PPF present little resemblance to the conventional polymers except for

NU
their organic nature [17]. Indeed, PPF are not characterized by the assembling of a repeating unit,

but by a random network presenting a cross-linking density significantly higher than


MA
conventional polymers (see Fig.1). In order to avoid confusion between plasma and conventional

polymers, the term “precursor” is sometimes preferred instead of monomer to name the molecule
D

from which the material is built. Nevertheless, both terms are accepted and currently employed in
TE

the plasma polymerization community.


P
CE
AC

Fig. 1. Schematic comparison of a plasma polymer film and a conventional polymer material

obtained from the same precursor/monomer.

4
ACCEPTED MANUSCRIPT

The formation of solid deposits from organic compounds using glow discharges is not new.

It was indeed first reported by Dutch researchers in 1796 [18]. These materials adhered tightly to

the walls of the glass-made reactors and were observed to be insoluble in most solvents.

PT
Nevertheless, they were considered as a nuisance until the work of Goodman who demonstrated

RI
that a 1 µm thick plasma-polymerized styrene film deposited on a titanium foil made a

satisfactory dielectric for a nuclear battery [19]. Since that time, the potential of these organic

SC
coatings has been revealed and a systematic investigation of the plasma polymerization process

NU
has been carried out. More information about the history of the plasma polymer science can be

found in Refs. [20-22].


MA
It is now demonstrated and accepted that PPF exhibit interesting physico-chemical properties
D

for organic materials such as high thermal, mechanical and chemical stabilities [4]. Moreover,
TE

due to their intrinsically good adhesion properties, numerous materials (e.g., glass, polymers,

metals), even with complex geometry (e.g., carbon nanotubes [23-26], micro/nanoparticles [27-
P
CE

31]), can be homogeneously covered [4]. All these features justify their use in a wide range of

applications. Historically, they were first developed in the search of physical barriers with
AC

applications in the field of corrosion protection [32, 33] and food packaging [34, 35]. In this

context, highly cross-linked PPF were needed. Such PPF are obtained when a high level of

precursor fragmentation occurs in the plasma. Thus, highly energetic conditions have usually

been employed due to the precursor fragmentation dependence on the energy dissipated in the

system [21]. As a result of these extensive fragmentation reactions, poor control of the PPF

chemistry was achieved.

Since the 1980s, with the rise of micro- and nano-technologies, plasma polymerization has

been further developed in the search for PPF with controlled and tailored chemistry while

5
ACCEPTED MANUSCRIPT

keeping their other inherent properties. In this context, functionalized PPF containing/supporting

–COOH [36-41], –OH [42, 43], –NH2 [44-51],, –COOR [52-56], –COR [57-59], –CFx [60-62], –

Br [20, 63], –SH [27, 64-68], thiophene-based units [69, 70] have been developed. The interest in

PT
this class of materials arises from their potential use in modern fields of applications including

RI
the development of supports for biomolecules immobilization [71-77] or cell growth [78-80],

interlayers for promoting adhesion of metal coatings [81], biocompatible [82] or antibacterial

SC
coatings [12, 83], controlled drug release coatings [84-87], super-hydrophobic surface [62],

NU
conductive layers [70],etc.
MA
For these applications, the chemical composition of the coatings is one of the most important

criteria defining its performances. Therefore, scientists have focused their efforts toward fine
D

control of the plasma polymer chemistry. It quickly appeared that this control can be obtained
TE

through the understanding of the PPF growth mechanism and more specifically of the phenomena

taking place at the plasma-substrate interface. Accordingly, investigating the plasma chemistry
P
CE

rapidly turned out to be a necessity. Numerous works have therefore been focused on the

investigation of the plasma phase during the PPF growth. Surprisingly, while other aspects of the
AC

field have been reviewed such as the growth mechanism of the layers [8, 21], their behavior in

liquid medium [82], their use for biological applications [72], their nanostructure [88] and their

surface analysis [89, 90], there are no documents summarizing the efforts made for a precise

evaluation of the plasma phase. Therefore, the present paper aims at reviewing the principal

works developed to evaluate the plasma chemistry during the low-pressure plasma

polymerization process. Particularly, we pay special attention to describe how these works have

contributed to enlarging the knowledge of plasma polymer growth at a molecular level.

6
ACCEPTED MANUSCRIPT

This review is organized as follows. In the first part, the plasma polymerization mechanism is

described. Then, an overview of the main achievements obtained by several research groups in

the field of plasma diagnostics related to the plasma polymerization process is presented. The

PT
most popular diagnostics methods employed for probing the species constituting the plasma are

RI
described, namely, mass spectrometry, gas phase Fourier transform infrared spectroscopy and

optical emission spectroscopy. The main results obtained by Langmuir and ionic probes to

SC
determine the plasma parameters and ion flux, respectively, are also discussed. In addition,

NU
throughout the paper, it is shown how theoretical calculations based upon the density functional

theory method have proven to be a powerful tool for assisting in the interpretation of the complex
MA
diagnostic data. Finally, conclusions and perspectives suggesting research strategies for

increasing the fundamental knowledge in the plasma polymer field are given.
D
TE

2. Plasma polymerization
P

In this section, we start with a brief summary of the main features of plasmas employed in
CE

the context of plasma polymerization. For a deeper description of plasma physics in surface
AC

processing, readers are invited to consult the review of Bogaerts et al. [91] and the book of

Lieberman and Lichtenberg [92]. Then, the main models developed in the literature for

describing the growth mechanism of plasma polymers are summarized. Finally, we discuss the

relationship between plasma parameters and film properties.

2.1 Plasma Fundamentals

A plasma is defined as a gas containing a mixture of electrons, ions, neutrals and photons. As

the electron and ion densities are equal, the plasma is macroscopically neutral. The plasma phase

was first described in 1920 by Irving Langmuir working on the development of vacuum tubes for

7
ACCEPTED MANUSCRIPT

large currents [11]. Although omnipresent in the universe and representing nearly 99% of the

matter (solar corona, solar wind, earth’s ionosphere), the natural presence of plasma on earth is

rare (e.g., lightning, aurora borealis) [93]. Nevertheless, the plasmas can be easily generated

PT
artificially via the electrical excitation of a gas, for example.

RI
The particles (electrons, ions and neutrals) constituting the plasma are in motion and undergo

SC
collisions during which energy is transferred. Each family of particle is often characterized by a

temperature related to its translational energy. In plasma science, the temperature of the particles

NU
is generally expressed using the eV as unit; the conversion factor is 1 eV = 11600 K. Depending
MA
upon potential differences in terms of temperature for the particles constituting the plasma, they

are classified in two main families: namely, at thermodynamic and at non-thermodynamic


D

equilibrium. For the former case, all the particles are characterized by a unique temperature,
TE

namely Te = Tn= Ti where Te, Ti and Tn represent the temperatures associated to electrons, ions

and neutrals species, respectively. In contrast, for non-equilibrium plasmas, the electron
P
CE

temperature strongly differs from that associated with the heavy particles, ions and neutrals: Te

>>> Ti ≥ Tn. Indeed, whereas the electron temperature typically ranges from 1 to 10 eV, the ion
AC

and neutral temperatures are close to room temperature, 0.025 eV (298K). This is why such

plasmas are often called “cold” plasmas. Indeed, although the temperature of the electrons is

high, their low density and heat capacity allow surfaces surrounding the plasma to remain at

relatively low temperatures [11]. This non-equilibrium feature is particularly attractive for

materials processing as the electrons can induce several chemical reactions without altering the

treated material by excessive temperature. Consequently, cold plasmas are predominantly used in

surface material processing including plasma polymerization [91]. For conventional plasma

sources (i.e., direct current, radio-frequency, microwave), the ionization degree is less than a few

8
ACCEPTED MANUSCRIPT

percent. Cold plasmas can be generated either at low (i.e., < 1 Torr) or high pressure, even at

atmospheric pressure. In the context of the plasma polymerization process, most studies are

conducted at low pressure and are the subject of this review. However, there is recently an

PT
increasing interest for operating the plasma polymerization process at atmospheric pressure.

RI
More details about this aspect can be found in the review of Merche et al. [94].

SC
Since the electrons are the energy vectors in the plasma, their density and energy are of

crucial importance for processes occurring in the gas phase. The electrons are assumed to be in

NU
thermal equilibrium at a given temperature Te [92]. Following kinetic gas theory, the electrons
MA
energy distribution function (EEDF) can be, to a first approximation, expressed by a Maxwell-

Boltzmann distribution as illustrated in Fig. 2 [20, 92]. Although in a conventional plasma


D

polymerization process, the majority of electrons possesses a kinetic energy centered around a
TE

given value (1-2 eV in Fig. 2), the distribution extends to high energy (from 4 to 20 eV). This

allows the occurrence of numerous chemical reactions in the plasma through collisional processes
P
CE

with electrons resulting in the formation of a great variety of species including radicals, ions and

photons. Since the degree of ionization in the plasma is often very low, the most probable
AC

collisions involving electrons occur with neutrals particles. The probability for a given reaction to

occur depends on the number of electrons able (from an energetic point of view) to induce the

chemical reaction. On this basis, considering organic discharges, in comparison to ionization,

dissociations reactions leading to the formation of radicals are more probable than ionization

events since most of the electrons in the plasma possess an energy of 1-4 eV similar to the energy

required to break simple organic chemical bonds (see Fig. 2); in contrast, the ionization energy of

organic molecules is approximately 10 eV. As a consequence, the concentration of radicals in the

plasma can be 103-105 higher than the ion density [22].

9
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
Fig. 2. Typical Maxwell-Boltzmann electron energy distribution function for low pressure
discharges. Adapted from [10, 20].

Another important aspect related to plasma technology is that owing to the significant
MA
difference in terms of thermal flow velocities between electrons and ions, a surface immersed in a

plasma naturally acquires a negative charge which acts to accelerate the positive ions toward the
D
TE

surface and repel most of the impinging electrons [8, 92]. For insulated surfaces (as generally

encountered for plasma polymer layers), at the equilibrium, the ion flow perfectly counterbalance
P

the electron flow leading to a stationary electrical potential named the floating potential (Vf). As
CE

a consequence of this phenomenon, near all surfaces including the substrate, regions called
AC

sheath characterized by a net positive charge develop. When ions enter the sheath, they are

accelerated and strike the surface with a kinetic energy ranging typically from 10 to 30 eV in the

case of a floating surface in the absence of collisions, i.e., when considering a low pressure

plasma polymerization process [38]. The influence of the bombarding ions on plasma

polymerization is obviously also crucial and is discussed below. It should be noted that when the

substrate is directly connected to an RF power supply, the negative self-bias developed at the

surface can be significantly higher than the floating potential and depends on the synthesis

10
ACCEPTED MANUSCRIPT

conditions [8]. Finally, another source of energy is also provided to the growing film due to the

irradiation of ultraviolet (UV) photons emitted by excited atoms/molecules.

2.2 Growth Mechanism

PT
Their unique properties make the PPF a specific class of materials arousing, over many

RI
decades, the curiosity of surface scientists who have developed strong efforts to understand their

SC
growth mechanism.

NU
The overall plasma polymerization mechanism involves both gas phase and surface reactions

[95]. The first step consists in the vaporization of an organic precursor in a deposition chamber.
MA
The activation takes place in the plasma phase through collisional processes between energetic

electrons and precursor molecules. As already mentioned, the most probable reaction consists of
D
TE

precursor dissociations reactions resulting in the formation of radicals. Historically, it was

therefore assumed that the growth of the layer mainly occurs through either radical-radical or
P

radical-molecule reactions. Indeed, the radicals possess an unpaired electron and therefore are
CE

highly reactive toward termination reactions with other radicals or toward addition reactions with
AC

unsaturated molecules (i.e., with double or triple bonds) [8].

Owing to the similarity between organic chemical bonds energy and the fact that the electron

energy is characterized by a distribution function, numerous fragmentation pathways are possible

[66]. As a consequence, a great variety of radicals is generally produced in the plasma whatever

the chemical nature of the organic precursor, hence contributing to the difficulty to obtain fine

control of the PPF chemistry.

Some of the processes involved in the synthesis of a PPF are summarized in the “Rapid Step

Growth Polymerization” (RSGP) model proposed by Yasuda in 1985. The model is built on the

11
ACCEPTED MANUSCRIPT

concept of recombination of reactive species and the subsequent reactivation of the resulting

products [96, 97]. The overall mechanism is schematically described in Fig. 3. Cycle 1 involves

the reaction of monoradicals whereas cycle 2 concerns biradical species, both originating from

PT
the interaction of the precursor with the plasma. Steps (1), (3), (4) and (5) correspond to addition

RI
reactions between the reactive species and (i) a stable molecule (which can be the precursor

molecule) containing a reactive site such as a double or triple bond (steps (1) and (4)) or (ii) other

SC
reactive species including radicals and biradicals (steps (3) and (5)). In both cases, the products

NU
formed can undergo other propagation reactions. It is obvious that the extent of addition reactions

in the gas phase depends on the pressure.


MA
In the plasma polymerization mechanism, the molecules formed through the recombination
D

between two radicals (step (2), termination reaction) can be reactivated via electron impact in
TE

contrast with the situation encountered in conventional polymerization. Consequently, the RSGP

mechanism can be viewed as a succession of termination reactions followed by the reactivation


P
CE

of the products. When a surface is exposed to such a plasma, a solid organic thin film arising

from the condensation of the reactive species produced in the plasma is deposited on the
AC

substrate. However, further “polymerization” reactions (initiation, addition, termination,

reactivation) can still occur at the plasma growing-film interface.

12
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D

Fig. 3. Schematic description of the Rapid Step-Growth Polymerization Mechanism. Adapted


from [96].
TE

In order to provide a more complete description of the plasma polymerization mechanism,


P

Yasuda added in his model the CAP (“Competitive ablation and polymerization”) principle,
CE

namely the simultaneous occurrence of film etching and deposition processes [21, 39, 96]. The
AC

synthesis of the coating thus results from a balance between both phenomena. Indeed, highly

reactive radicals (e.g., .O., .F, .S., etc) can be produced and react at the interface to form stable

molecules (e.g., CO2, CO, CS2, CF4). The products may desorb from the growing film leading to

the removal of part of the coating (ablation). These stable molecules cannot take part in the

growth of the film anymore and are either (i) pumped out of the reactor or (ii) reactivated through

electron collisions. The occurrence of such surface reactions directly depends on the amount of

energy provided to the growing film interface by ion bombardment and UV photon irradiation. It

should be noticed that these stable molecules can also be formed via gas phase reactions reducing

13
ACCEPTED MANUSCRIPT

the amount of reactive species which can potentially take part in layer growth. The latter process

is named the “scavenger effect” and can also affect the PPF chemistry [40].

At this stage of the discussion, owing to their relative low abundance in the discharge

PT
compared to radicals and neutrals, the influence of ions in the growth mechanism has been totally

RI
excluded. Nevertheless, as already mentioned, in plasma processing, a surface exposed to the

SC
discharge is continuously submitted to a flow of ions for which the typical kinetic energy ranges

from 10 to 30 eV for a floating surface. This supply of energy is enough to induce chemical bond

NU
breaking leading to the formation of surface dangling bonds which can act as preferential
MA
adsorption sites for reactive species in the plasma [61, 98, 99]. The impact of this phenomenon is

taken into account in the ion-Activated Growth Model (AGM) proposed by d’Agostino in which
D

the formation of surface defects through ion interaction is considered [60, 98]. Based on the
TE

AGM, the precursor can be incorporated in the growing film via a surface reaction with a radical

site through for example the opening of a double bond. Such reaction is referred to induced
P
CE

plasma polymerization in contrast to plasma-state polymerization for which the activation of the

molecule in the plasma is an essential step. Initially, the AGM was developed for the growth of
AC

fluorine-based coatings, but it can also be applied to other PPF families [61]. In addition, the

impinging ions can also be responsible for other phenomena such as ion-assisted etching or

coating densification as it will be described later. For sake of completeness, it has to be

mentioned that surface reactive sites can also be created through UV photons irradiation of the

growing film interface.

While considering ions as active species defining the PPF properties, the AGM still implies

that the density of ions is so low that any mass deposited by ions would be insignificant [100].

However, experimental evidence has clearly revealed that the ions play a greater role in the

14
ACCEPTED MANUSCRIPT

formation of a PPF [100]. Indeed, recent studies pointed out that under certain experimental

conditions, the contribution of condensing ions cannot be neglected [8, 57, 58, 101-104]. In some

cases, some authors even claim that the ions are the main species responsible for PPF growth

PT
[57]. The origin of the greater role played by the ions in PPF formation is justified by several

RI
factors. For instance, the presence of a sheath at the plasma growing film interface accelerates the

ions toward the film surface, significantly reducing the neutral-to-ion flux ratio at the surface in

SC
comparison with that measured in the gas phase [8, 102]. Considering a low pressure plasma,

NU
e.g., 1 mTorr, the surface flux ratio of neutrals to ions is estimated to be approximately 20 times

lower than the corresponding density ratio in the plasma [102]. Another important factor to take
MA
into account is the sticking probability of the ions. From studies investigating layers grown by

hyperthermal ions, a sticking coefficient of 0.2-0.5 can be estimated for the ionic species
D

depending on their energy [105, 106]. On the other hand, for radical-based molecules, the
TE

sticking coefficient has been estimated to vary over a wide range (~10-4 - ~1) depending on the
P

chemical nature of the radical (e.g., unsaturation degree) [107-109] and the activation of the
CE

surface through the formation of open bond sites at the interface [99, 107, 110, 111]. Therefore,
AC

as integrated in the AGM, a synergetic effect between ions and neutrals has to be considered

since increasing the ion flux enables the formation of a higher number of reactive sites at the

surface, thus promoting the grafting of radicals.

To summarize, although nowadays it is accepted that the contribution of ions in the deposited

film is significant, their exact role in the plasma polymerization mechanism is still debated [112,

113]. It is, however, important to shed light on their impact for defining PPF properties as

recently demonstrated by Michelmore et al. for the mechanical properties of the deposited layers

[58].

15
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
Fig. 4. Overall mechanism of plasma polymerization.
D

The combination of the different models described in this section (schematically represented
TE

in Fig. 4) allows us to provide an overview of the main reactions taking place in the plasma and

at the PPF surface. This complex mechanism is at the origin of the poor control of the chemistry
P
CE

of such thin films. Indeed, it has been extensively reported that functionalized PPFs contain

numerous chemical functions even if the organic precursor is monofunctional [21, 67]. This
AC

complex mechanism also explains the high branched and cross-linking degree of plasma

polymers triggering their attractive physico-chemical properties. This irregular structure results

from the random recombination of the numerous plasma-generated fragments followed by

successive rearrangements, fragmentations, reinitiations reactions, etc.

Another consequence of this complex mechanism is the presence of radicals in the PPF

network after synthesis [72, 114-118]. These radicals have been identified to be at the origin of

the well-known “aging” of PPFs, namely oxidation of the material [49, 82, 114, 119].

Nevertheless, it has been recently reported that they can be advantageously exploited for inducing

16
ACCEPTED MANUSCRIPT

a radical conventional polymerization [117, 120-124]. The trapped radicals can also serve as

reactive sites for the grafting of proteins [125, 126]. Another aging phenomenon is related to the

potential presence of embedded molecules in the plasma polymer matrix [49, 82, 127] which can

PT
be released in air during storage of the material [66, 128-132] or after immersion of the coatings

RI
in a liquid medium [36, 49, 68, 133-135]. The latter could be detrimental for biological

applications requiring the immersion of the layers in solution [82, 127, 136, 137].

SC
2.3 Control of PPF properties through deposition parameters

NU
The chemical composition of the layers (e.g., the density of a particular chemical function) as
MA
well as their cross-linking degree are strongly affected by the deposition parameters: the plasma

source (e.g., radiofrequency, microwave,…), the power dissipated in the discharge, the working
D

pressure, the precursor flow rate, the substrate temperature, etc. [4, 8, 17, 21, 22, 131, 138-140].
TE

In addition, some geometric parameters related to the design of the chamber also affect the
P

coating properties. Examples include the location of the precursor inlet, the distance between the
CE

plasma source and the substrate, etc. The dependence of the process with the reactor geometry

makes therefore difficult to compare the “same” PPF from one deposition chamber to another
AC

[22, 141].

The need for PPFs with a precise chemical composition and structure requires a clear

understanding of the relationship between the process parameters and PPF properties. Very often,

an empirical approach is utilized and the common strategy consists in separately studying the

influence of one process parameter, while keeping the other ones constant.

For a given reactor configuration, the influence of the operational parameters can be

understood by considering their impact on the plasma characteristics. For example, the applied

17
ACCEPTED MANUSCRIPT

power alters the electron energy distribution function and density which govern the nature and

the rate of the chemicals reactions occurring in the plasma [142]. This, in turn, influences the

chemistry of the coatings and their physico-chemical properties.

PT
According to the plasma polymerization mechanisms, the most important factor affecting the

RI
properties of a PPF is the energy applied per molecule which governs the degree of precursor

SC
fragmentation in the discharge [96]. In this context, Yasuda proposed a composite parameter (the

so-called Yasuda parameter): W/FM where W, F and M are the power dissipated in the discharge

NU
in J/s, the monomer flow rate in mol/s and the molecular weight of the monomer in kg/mol,
MA
respectively [97]. The Yasuda parameter represents the energy input per unit mass of monomer.

The term W directly scales with the electron density and hence governs the collision frequency
D

between electrons and precursor molecules [40, 95]. For a fixed working pressure, the term 1/F
TE

scales with the residence time of the particles in the discharge, thus also influencing the extent of

precursor fragmentation. Therefore, the concentration of activated species in the plasma directly
P
CE

depends on the W/FM parameter. Regarding the evolution of the deposition rate with W/FM, two

principal domains of plasma polymerization can be identified (see Fig. 5). At low W/FM, the
AC

deposition rate evolves linearly with W/FM. In this region, the so-called “energy-deficient

domain”, ample monomer is available and the supply of energy is the limiting factor for

increasing the deposition rate. In this domain, increasing the supply of energy in the discharge

results in an increase in the concentration of film-forming species through collisional processes.

Above a critical W/FM value, the deposition rate becomes constant since the precursor

fragmentation has attained its maximum. This regime is called “monomer-deficient” as sufficient

energy is provided, but the precursor feed rate into the chamber is the limiting factor [143]. In

addition, under high energy conditions, ion-induced etching reactions can be favored and this

18
ACCEPTED MANUSCRIPT

leads to a decrease in the deposition rate. The extent of ablation phenomenon strongly depends on

the chemical composition of the discharge. It is, for example, well known that fluorine-based

discharges are quite sensitive to ablation reactions [4, 61].

PT
Complementary to the Yasuda concept and as a further development, Hegemann et al.

RI
developed a macroscopic approach for describing the plasma polymerization process based on

SC
the concept of chemical quasi-equilibrium [39, 40, 95, 144-147]. In this case, the plasma is

divided into (i) an active zone in which the activation takes place through collisional processes

NU
leading to the production of film-forming species and (ii) a passive zone where the deposit is
MA
formed. This approach states that the mass deposition rate Rm (expressed in g/cm2s ) depends on

the parameter W/F following a quasi-Arrhenius behavior according to Eq.1 [144]:


D

Rm E
= G exp(- a ) , (1)
TE

F W/F
P

where G is a reactor and process dependent factor related to the maximum monomer conversion
CE

into film growth and Ea represents an apparent activation energy related to the PPF system. It

should be emphasized that the W/F parameter can be adjusted depending on the reactor
AC

configuration (e.g., symmetric vs. asymmetric), thus making the macroscopic approach suitable

for facilitating the transfer of a plasma polymerization system from one chamber to another [146,

148, 149].

A typical example of an Arrhenius-type plot (ln(Rm/F) versus the inverse specific energy,

(W/F)-1, is shown in Fig. 6 for the plasma polymerization of methane. Ln (Rm/F) evolves linearly

with (W/F)-1 pointing to an Arrhenius-like behavior over the range of W/F investigated. In this

regime, the film grows mainly through radical reactions and the kinetic limiting factor is the

19
ACCEPTED MANUSCRIPT

production rate of the film-forming species in the plasma. It is assumed here that the

fragmentation pattern of the precursor in the plasma is identical with W/F and that the formation

rate of radicals increases exponentially with energy input [148]. It has been reported that Eq. 1

PT
holds for many monomers and gas mixtures for a certain range of W/F [146, 147, 150]. In Fig. 6,

RI
from the negative slope of the linear fit, Ea, which is related to an ensemble of fragmentation

reactions in the active zone, can be deduced [151]. Its value is monomer specific and correlates

SC
with the bond dissociation energy of the precursor [146]. Compared to the Yasuda approach, the

NU
activation energy separates the plasma polymerization into the energy and monomer deficient

regime [144, 148].


MA
When applied to several PPF families for a broad range of energy conditions, the
D

macroscopic approach has revealed that more than two deposition regimes (i.e., the energy and
TE

monomer deficient regime at high and low energy levels, respectively) have to be considered

depending on the plasma polymerization system. This could be related to different growth
P
CE

mechanisms. For instance, deviations from Eq.1 at low energy levels could indicate a

predominant ionic oligomerization mechanism and/or the grafting of intact monomers on reactive
AC

sites at the interface [148]. At high energy levels, the drop in the deposition rate observed for

instance in the case of nitrogen- and oxygen-based gas mixtures could provide information about

variations in plasma chemical pathway mechanism or ion-induced etching reactions [40].

Although the macroscopic approach has revealed its potential for describing the plasma

polymerization process, defining the plasma polymer formation mechanism based on the

evolution of the Rm/F as a function of (W/F)-1 in an Arrhenius-type plot is not straightforward and

could lead to erroneous conclusions without additional data from, for example, plasma diagnostic

20
ACCEPTED MANUSCRIPT

measurements. This has led to a very interesting debate in the plasma polymer community

regarding the suitability of the macroscopic approach [39, 148, 151-154].

The energy applied to the plasma polymerization process affects not only the deposition rate,

PT
but also the composition and the cross-linking degree of the layers. At high power, more cross-

RI
linked PPFs are formed because of the extensive fragmentation of the precursor yielding a higher

SC
quantity of small-molecular weight film-forming radicals. At the same time, ion bombardment

becomes more intense and may also contribute to the densification and cross-linking of the

NU
growing film (see below). These high energy conditions are suitable for obtaining good barrier
MA
properties finding applications in food packaging and corrosion protection [155]. On the other

hand, plasma polymerization conducted at low energy conditions provides a low degree of
D

precursor fragmentation and a high retention of the functional group hosted by the precursor. This
TE

can be explained by the small amount of electron-induced collisions with precursor molecules

leading therefore to an activation process rather than a complete disintegration of the initial
P
CE

chemical structure. This mode of operation is therefore more suitable for biological applications

[8, 21, 41, 47, 48, 55, 72, 82, 156].


AC

Fig. 5. Typical evolution of the deposition rate as a function of the Yasuda parameter, W/FM
illustrating the different regimes of plasma polymerization [143].

21
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 6. Arrhenius-like plot of the deposition rate versus the inverse energy input for a methane
discharge. Adapted from[146].

NU
MA
D
P TE
CE
AC

Fig.7. (a) Plasma polymer film (from C2H4, NH3/C2H4 and CO2/C2H4) density vs. the momentum
density during film growth. The linear evolution indicates densification by momentum transfer.
(b) Chemical composition of a plasma polymer from NH3/C2H4 and CO2/C2H4 depending on the
momentum density during film growth. The filled symbols indicate the relative amount of
oxygen and nitrogen. The open symbols show the functional group density which is reduced with
the increasing densification. Adapted from [157].

22
ACCEPTED MANUSCRIPT

As already mentioned, the energetic conditions at the growing film surface due to ionic

bombardment are a key factor controlling film properties. Therefore, in complement to the

macroscopic approach, with the aim to rationalize the influence of the ion bombardment in

PT
plasma polymerization based on the synthesis conditions, Hegemann et al. have introduced a new

concept, namely the momentum density ( πsurf ) defined as the momentum flux per deposition rate

RI
SC
following Eq. 2 [157]:

Γi E mean

NU
πsurf = 2mi , (2)
R
MA
where mi,  i , Emean are the average mass, the flow and the mean energy of bombarding ions,
respectively. R represents the deposition rate.

For example, the authors reported a linear correlation between πsurf and the density of the
D
TE

coatings (directly related to the cross-linking degree) for plasma polymerization from discharges

of pure C2H4 as well as NH3/C2H4 and CO2/C2H4 mixtures (Fig. 7). This trend is explained by the
P
CE

increase in chemical bond breaking at the surface and subsequent random recombination of

radicals. At the same time, although the N/C and O/C ratios are constant for a certain range of
AC

energetic conditions, the NH2/N and COOH/COOR ratios scale inversely with densification as

induced by πsurf . This example illustrates the important role played by the bombarding ions in the

plasma polymerization process as reported recently for the control of coating stability in aqueous

medium [50, 158].

In view of the modern applications of this class of coatings, especially in the field of

biotechnology, the control of the chemical composition of the PPF has become an aspect of

increasing importance. In this context, the optimization of “conventional” plasma parameters

(mainly the W/FM parameter) presents some limitations. Indeed, even when using low energy

23
ACCEPTED MANUSCRIPT

conditions, an irregular structure predominates and the density of the targeted function remains

low. With this problem in mind, a breakthrough appeared when the pulsed plasma technique,

reported in the field of plasma polymerization in 1972 by Tiller et al., was introduced [21]. The

PT
idea was to further reduce the extent of fragmentation of the precursor and hence the side

RI
reactions accounting for the formation of other functionalities than the one hosted by the

precursor. Today, pulsed plasma polymerization has become a well-established method for the

SC
synthesis of functionalized PPF [37, 41, 42, 44, 46, 48, 54, 59, 68, 70, 159-161].

NU
The pulsed approach consists in producing the discharge intermittently according to the pulse
MA
frequency. The mean power dissipated into the discharge can be easily modulated by adjusting

the time during which the plasma is switched ON (ton) and OFF (toff) [47].
D

During the plasma “ON” time, electrons, ions, radicals and photons are produced. The ton
TE

period can be approximated to a discharge operating in Continuous Wave (CW) mode sustained
P

at a power equivalent to the power applied during the plasma “ON” time [37]. During this period,
CE

the layer grows through the complex mechanism previously described. When the plasma is

switched OFF, due to the recombination of the electrons and ions at the reactor walls, the electron
AC

density as well as the floating potential naturally developed at the substrate rapidly decrease as

experimentally measured [41]. Consequently, in this case, ion bombardment and photon

irradiation rapidly disappear, limiting the side reactions at the surface and thus favouring the

reaction of radicals present in the gas volume with the nucleation sites generated during ton.

Depending on the chemistry of the starting molecule, the precursor itself can participate in

addition reactions with the radicals sites present at the surface of the growing film even during

toff. This has been experimentally demonstrated for the pulsed plasma polymerization of acrylic

24
ACCEPTED MANUSCRIPT

acid [37]. This precursor, containing a double bond, can therefore easily undergo a propagation

reaction through a radical mechanism involving the opening of the double bond. In this case, the

pulsed mode can ideally be viewed as a succession of activation reactions (ton) followed by

PT
propagations steps during toff, thus favouring the incorporation of the targeted chemical group in

RI
the coatings [45].

SC
If saturated molecules are considered, growth events during toff can be excluded. Indeed, in

this case, during toff, the radicals rapidly recombine at the interface and the precursor itself does

NU
not participate in the formation of the film, as revealed for the pulsed plasma polymerization of 3-
MA
fluoroaniline [160]. Nevertheless, even in this case, the use of the pulsed plasma polymerization

strategy is beneficial as mean powers lower than those required to sustain the discharge in the
D

CW mode can be employed, thus allowing a significant reduction in the degree of fragmentation
TE

of the precursor and in turn enhancing the retention of the chemical group of interest [37, 46].
P
CE
AC

25
ACCEPTED MANUSCRIPT

3. Which tools can we use to probe the plasma polymerization process?

As detailed in the previous section, in plasma polymerization, the numerous species present

in the discharge (i.e., electrons, ions, radicals, stable molecules and photons) react with each

PT
other and with the growing film via a multitude of interaction pathways [117]. This complex

RI
mechanism makes the assessment of each individual reaction as well as the prediction of the

SC
coating properties very challenging. Furthermore, depending on the chemical nature of the

precursor and the synthesis conditions, especially at low energy conditions, some specific

NU
reactions can predominate. This would also influence the final features of the formed layers.
MA
Therefore, it has rapidly become obvious that for obtaining a good control of the PPF chemistry,

a knowledge of how a PPF grows at a molecular level is essential [8].


D

3.1 Diagnostic techniques


TE

When concluding one of his papers dealing with the development of the macroscopic
P

approach, Hegemann wrote: “ It is expected that further clarification of the complex processes
CE

taking place during plasma polymerization can be achieved by the use of plasma diagnostics in
AC

combination with the macroscopic approach” [146]. Therefore, in addition to the development of

empirical models (e.g., Yasuda, Hegemann), the plasma polymer community has more and more

attempted to characterize the process by employing state-of-the-art plasma analysis (mass

spectrometry, gas phase Fourier Transformed Infrared Spectroscopy, optical emission

spectroscopy, electrostatic and ion probes) and surface analysis (X-ray Photoelectron

Spectroscopy-XPS, Time of Flight Secondary Ion Mass Spectrometry-ToF-SIMS) tools.

Several diagnostic methods have been developed and employed during the last 30 years for

detailed comprehensive study of the plasma chemistry. For evaluating the chemical composition

26
ACCEPTED MANUSCRIPT

of the plasma, non-intrusive optical diagnostics methods including optical emission spectroscopy

(OES) and gas-phase Fourier Transform Infrared Spectroscopy (GFTIR) as well as mass

spectrometry (MS) have been used extensively. In complement to these techniques, ion probes,

PT
especially designed for operating in organic discharges, have also been developed with the aim to

RI
measure absolute ion fluxes toward surfaces facing the plasma. Finally, governing the production

rate of species in the plasma, the electron density and temperature can be evaluated using

SC
electrostatic probes. Table 1 summarizes the main features of the diagnostic methods employed

NU
in the context of low-pressure plasma polymerization. The techniques are described in more

detail in the next section of this review.


MA
D
P TE
CE
AC

27
ACCEPTED MANUSCRIPT

Table 1. Main features of the plasma diagnostics methods described in this review. The last

column refers to the relevance of the diagnostic tool as a plasma polymerization diagnostic: + =

low relevance tool, ++ = relevant tool, +++ high relevance tool.

PT
Relevance in
view of the

RI
Species probed or Time elucidation of
Method Comments
measurable parameters -resolved the growth

SC
mechanism of
PPF
Electron

NU
Electrostatic or Yes
density/temperature,
Langmuir [41, 162, +
plasma/floating Potential
probe 163]
[142, 162] MA
Semi-
quantitative in
Excited neutrals/ ionic
Yes some cases if
OES species [69, 70, 132, 156, ++
[166] using an internal
164, 165]
D

standard [156,
166-168]
TE

-Semi-
- Neutrals and ions
quantitative [48,
depending on the analysis
54, 66, 68, 131,
P

mode [37, 38, 54, 57, 58,


Yes for 172]
Mass 66, 68, 101, 128, 169, 170]
CE

ions [41, +++


Spectrometry
163] -Quantitative
-Ion energy distribution
for neutrals in
function of ions [38, 104,
AC

some particular
171]
cases [173-175]
Absolute ion flux
Ion Probe toward a surface [10, 101, No ++
102, 104, 176]
Quantitative
Vibration frequency of using a
GFTIR chemical bonds [52, 99, No calibration +++
177, 178] procedure [52,
99]

28
ACCEPTED MANUSCRIPT

3.2 Theoretical support for a better description of the data

If the evaluation of the plasma chemistry is a challenging task, the interpretation of the

accumulated data is often, at least, as difficult. This is why the use of theoretical tools is relevant

PT
in this context when used in close synergy with experimental measurements [179]. Modelling the

RI
plasma chemistry is a formidable task in view of the large number of species involved and the

SC
large diversity of intervening processes (interaction between light and plasma species, heat

propagation, chemical reactions, thermal diffusion of the plasma species, etc.). Many theoretical

NU
studies rely on mechanistic models involving generally only a fraction of the processes and/or
MA
species, without any atomistic detail; in practice, kinetic equations are solved to describe, as a

function of time, for instance the generation of the different species in the plasma (together with
D

their degree of charging) and their spatial distribution. The main limitation of such approaches is
TE

that the results heavily depend on the actual values chosen for the kinetic rates, in particular for

the underlying activation energies. Those rates can be inferred from experimental measurements
P
CE

or estimated from sophisticated quantum-chemical calculations. A critical comparison between

reference data and the results of the simulations is required to validate the chosen parameters. A
AC

refinement is to account for the chemical structures by using molecular dynamics simulations

based on force fields (i.e., expressions yielding the relative energies of a given system in different

geometries based on a series of bonded and non-bonded energetic terms). These methods cannot

account for electronic excitations of the species or for the presence of free electrons; moreover,

although standard force fields cannot describe the formation of chemical bonds, the

implementation of reactive force fields (such as ReaXX force fields [180]) allows one to

circumvent this limitation. Such simulations are deterministic in the sense that they can describe

the trajectory of all species. There are, however, two main limitations associated with such

29
ACCEPTED MANUSCRIPT

simulations: (i) the quality of the results critically depends on the nature of the interatomic

potentials used to describe the van der Waals interactions and the charge assignment on the atoms

to evaluate the Coulomb energies; (ii) those simulations cannot be solved analytically and require

PT
a time discretization, with a very small time step (of the order of a femtosecond). Accordingly,

RI
the simulations are typically run over a limited time (typically a few hundred nanoseconds) and

cannot account for slow kinetic processes. However, exploration of the conformational space can

SC
be accelerated by increasing the temperature or by coupling the force field to Monte Carlo

NU
simulations based on a random (compared to thermodynamic) sampling of the system.
MA
A theoretical approach much less adopted in the field of plasma chemistry is to perform

quantum-chemical calculations to shed light on important properties of the plasma chemistry. The
D

most popular method used is Density Functional Theory (DFT) which limits the size of the
TE

system to be considered (typically up to one hundred heavy atoms), but provides key information

such as the free enthalpy of chemical reactions or activation energies. DFT is not a universal
P
CE

theory and actually comes with many different flavors based on the choice of the functional and

basis set. It is thus always recommended to validate the selected DFT approach by comparison to
AC

experimental data or highly sophisticated Hartree-Fock based calculations. DFT can also be

coupled to molecular dynamics simulations, for instance using the Car-Parrinello method though

at a much more extensive computational cost compared to force-field calculations [181]. DFT

typically sheds light on a very specific aspect in the plasma, generally to assess the change in the

electronic energy (or in the enthalpy or free enthalpy) associated with a reaction and activation

energies related to computed reaction pathways. For illustration, DFT has been used to model the

growth of CVD diamond, showing the different steps allowing for the insertion of an additional

CH2 group on the surface starting from a CH3 radical [182]; in this study, DFT was actually

30
ACCEPTED MANUSCRIPT

coupled with a force field (within a Quantum Mechanics/Molecular Mechanics – QM/MM -

approach) in order to treat the bulk diamond material at a lower level of theory to account for the

electrostatic environment in the simulations and the part of the surface where the grafting occurs

PT
with DFT to access electronic properties. Another study reported the use of DFT calculations to

RI
select the best precursors for the plasma polymerization of organosilicates by computing bond

dissociation energies and free enthalpies of reactions [183]. Interestingly, algorithms to find

SC
transition states (and hence to estimate activation energies) are now implemented in many

NU
quantum-chemical packages, using, for example the intrinsic coordinate reaction (IRC) theory

[184]. The transition state can be validated by: (i) a frequency analysis showing a negative
MA
frequency for one mode, and (ii) by a steepest descent algorithm from the transition state

ensuring that the correct reactants and products are reached. This approach has been exploited in
D

a study modelling the production of ethane from methane in a plasma [185]. Interestingly, all
TE

these studies involved the B3LYP functional of DFT owing to its good performance in
P

reproducing experimental enthalpies of formation [186].


CE

Based on the previous considerations, the DFT method appears to be promising for a detailed
AC

description of the chemical reactions taking place in the plasma. For a decade, our group has

developed several strategies using DFT calculations in combination with standard diagnostic

techniques to enhance the understanding of plasma chemistry, as it will be exemplified in the

next section.

31
ACCEPTED MANUSCRIPT

4. Overview of the diagnostics methods

In this section, each diagnostic technique is individually presented discussing its advantages

and drawbacks. After a brief description of the basic principle of the method, particular attention

PT
is devoted to the link between diagnostic data and growth mechanisms. Furthermore, in several

RI
examples, we also show how the use of DFT calculations can be employed for assisting and

SC
completing the complex diagnostic data.

NU
4.1 Electrostatic probe

The electrons are the energy vehicles in the plasma and thus govern the production rate of
MA
reactive species through the dissociation/ionization of the precursor. Therefore, the knowledge of

the electron temperature, density and electron energy distribution function (EEDF) are mandatory
D
TE

for clarifying chemical reactions pathways. These plasma parameters, together with the floating

potential, can be obtained by using the well-known electrostatic probe (also referred to as a
P

Langmuir probes).
CE

This method is relatively easy to implement during PECVD experiments. The probe itself
AC

typically consists of a tungsten wire, which is biased with respect to the ground potential by a DC

generator. By sweeping the voltage from negative to positive values, e.g., from – 50 V to + 50 V,

the probe surface first collects positive ions (regardless of their charge state or their chemical

nature) and, as the voltage becomes positive, plasma electrons. By processing the current-voltage

characteristics, one can obtain plasma parameters as a function of process parameters. More

information on the theory as well as technical aspects of probe diagnostics can be found in Refs.

[92, 187].

32
ACCEPTED MANUSCRIPT

Langmuir probes have been used for a long time to analyze deposition plasmas whatever the

process, i.e., Physical Vapor Deposition or PECVD. However, in each case, scientists have to

face the same problems, namely: the deposition of the coating onto the probe surface and the

PT
probe body and the fact that the measurement is spatially localized. The coating deposition issue

RI
might be somehow exacerbated during the deposition of plasma polymers owing to the relatively

large deposition rate (as compared to, e.g., magnetron sputter deposition). Obviously, researchers

SC
have devised and upgraded their tools in order to minimize coating deposition related issues, e.g.,

NU
by programming cleaning routines between data acquisition steps. The later can be achieved by

positively biasing the probe for several seconds.


MA
Biederman et al. used a Langmuir probe during the DC deposition of hydrophilic films using
D

hexane/Ar/H2O mixtures [188]. Fig. 8 shows how the (absolute) probe current varies as a
TE

function of the probe voltage for various gas mixtures.


P
CE
AC

Fig.8. Evolution of the probe characteristic for Ar/H2O and Ar/H2O/hexane gas mixtures [188].

From their measurements, one can learn that the electron population is divided into two

groups, slow and fast electrons, when the plasma is ignited in pure argon gas (24 Pa). The low

33
ACCEPTED MANUSCRIPT

energy group is characterized by a Maxwellian–like EEDF. By the addition of water in the Ar

plasma, the electron temperature decreases from 0.70 to 0.15 eV, and the EEDF is no longer

Maxwellian. The electron density, the floating and plasma potentials also change. The electron

PT
density (ne) varied from ~3x1015 m-3 to less than ~1x1015 m-3 as the percentage of H2O was

RI
increased from 0 to 50%. The floating potential has a relatively constant value around 2.5 V as

the H2O was varied in the same proportion. However, the plasma parameters changed

SC
dramatically as hexane was added to the mixture; the electron temperature decreased to less than

NU
0.1 eV and the plasma and floating potentials varied by more than 20 V. To summarize these

observations, one can say that a small addition of hexane dramatically modifies the properties of
MA
the plasma. Obviously, these modifications will change how the plasma species interact with the

chamber walls and the growing plasma polymer film.


D
TE

Time-resolved Langmuir probe studies were also reported for acrylic acid pulsed discharges

in order to determine the temporal evolution of the density and temperature of the electrons as
P
CE

well as the plasma potential adjacent to the deposited film [41, 162]. An example of such a study

is depicted in Fig. 9. In this work, the probe was either compensated or uncompensated. Most of
AC

the time, when used to analyze RF plasmas, Langmuir probes must be compensated by adding an

RC circuit to the probe circuit. This modification allows for filtering the RF harmonics, which

may lead to a distortion of the I-V characteristics. With such a filter, I-V curves can be processed

as if they were obtained in DC plasma.

From Fig. 9, it can be learned that in such a pulsed discharge, the plasma parameters are

time-dependent. Electron densities and temperatures peak during the ON-time, when the

electrical energy is transferred to the plasma species. The electron temperature and density then

quickly decay as the power supply is switched off. As a consequence, the gas phase chemistry,

34
ACCEPTED MANUSCRIPT

which is mainly promoted through electron impact, can be modulated by varying the duty cycle

(ton/toff + ton). From Fig. 9, it can also be seen that the electron density and temperature increase

when the power applied to the plasma is increased. The pulse amplitude and the duty cycle are

PT
thus key parameters when devising the synthesis of plasma polymer films through PECVD

RI
processes.

SC
NU
MA
D
TE

Fig. 9. Evolution of (a) the electron density and (b) electron temperature during the pulse period

for plasma polymerization of acrylic acid. In each case, the pulse frequency equals 500 Hz and
P
CE

the duty cycle is set to 50%. Black circles, triangles and lozenges correspond to a peak power of

50, 40 and 20 W, respectively. Adapted from [162].


AC

In order to minimize the probe coating deposition issue, a modification of the classic probe

setup in which a loop of fine wire is heated has been implemented [142]. Such probes are also

referred to as emissive probes because biasing the wire with DC or AC induces strong electron

emission from the probe itself. More details can be found in Ref. [142].

For sake of completeness, it has to be mentioned that diagnostic techniques other than

Langmuir probes are available to obtain relevant information on the electron population. As an

example, Guimond et al. used microwave interferometry to estimate the electron density in

35
ACCEPTED MANUSCRIPT

organic plasmas [95]. More information on this interferometry technique can be found in Ref.

[189]. This method offers some advantages as compared to the more conventional electrostatic

probes, namely (i) the measurement is not perturbed by film deposition and (ii) it is not necessary

PT
to take into account the ion composition in the plasma sheath in order to deduce the plasma

RI
electron density of the plasma. The main disadvantage lies in the lack of information about the

electron temperature.

SC
Using microwave interferometry, Guimond et al. measured more than a ten-fold increase in

NU
electron density, which increased from ~ 2.5x1015 m-3 to 3.5x1015 m-3 as the power was increased
MA
from 5 to 100 W in C2H4 and NH3/C2H4 discharges [95]. It should be noted that the evolution of

the electron density saturated as the power delivered to the plasma reached ~60 W. Gas heating is
D

invoked to explain this behavior. Similar trends in the electron density vs applied power were
TE

obtained by the same authors in C2H4 and C2H4/CO2 discharges [39]. These experimental results

were used to further describe the growth mechanism and to distinguish the relative influence of
P
CE

both chemical and physical plasma processes.

In conclusion, Te, ne and EEDF are key data that should be known in order to understand the
AC

fragmentation and excitation/ionization reactions taking place in the plasma, and in fine, to better

understand the mechanistic formation of the film. Electrostatic probes represent a simple and

rather cheap way to obtain this information. However, research scientists must keep in mind the

limitations of the technique such as the covering of the probe issue. In order to overcome these

limitations, more sophisticated, non-intrusive diagnostic methods can be implemented, such as

the microwave interferometry.

36
ACCEPTED MANUSCRIPT

4.2 Optical Emission Spectroscopy

Investigating plasma chemistry is essential for identifying reactive species taking part in PPF

formation, and hence controlling film properties. Among other techniques (i.e., mass

PT
spectrometry and the gas phase Fourier transform infrared spectroscopy detailed in other

RI
sections), this can be achieved by optical emission spectroscopy (OES) which is based on the

SC
collection, by a spectrometer, of the radiation coming from the plasma. Optical emission from a

plasma occurs primarily through electron impact excitation of atoms or molecules according to

NU
Eq. 3:
MA
N + e-  N* + e- , (3)

where N* is the excited state of the species N.


D
TE

This is followed by a relaxation to a lower energy state releasing a photon of energy equal to
P

the difference between the two energy states (Eq. 4):


CE

N*  N + hν , (4)
AC

where h is Planck constant and ν is the frequency of the emitted photon.

The emission of specific frequencies can be used to identify the species present in the

studied plasma. More details about the technique can be found in the following Ref. [92, 187,

190, 191].

OES studies are well documented in the literature related to low pressure sputtering plasmas

such as magnetron discharges because, in this situation, the spectra are usually sparse and

relatively easy to analyze. Indeed, most of the excited species inside such plasmas are sputtered

metal atoms and argon atoms. Diatomic molecules such as oxygen or nitrogen are added to the

37
ACCEPTED MANUSCRIPT

argon background gas if oxide or metal nitride compounds have to be synthesized. In the case of

plasma polymerization, which makes use of more complex organic molecules, the optical spectra

becomes crowded by emission bands originating from the multiple electronic, vibrational, and

PT
rotational excited states of the precursors. Fig. 10 shows a typical emission spectrum of

RI
acetone/CO2 plasma.

SC
NU
MA
D
P TE

Fig.10. Emission spectrum recorded in the ultraviolet-visible range during the plasma
CE

polymerization of acetone/CO2 [164].


AC

Although, numerous emission bands and lines overlap and are poorly, or not, resolved, some

lines can be unambiguously identified such as Hα, OH, CH, CO, CO2 and N. The presence of

nitrogen emission might originate from impurities incorporated during the deposition process.

The presence of the other species result from the dissociation/rearrangement reactions taking

place in the plasma, hence highlighting the complex gas-phase chemistry in this kind of plasma.

Therefore, OES allows for a qualitative description of the plasma chemistry enabling the

investigation of the chemical reaction pathways occurring in the discharge as reported for

38
ACCEPTED MANUSCRIPT

nitrogen [23, 192-195], fluorine [196, 197], sulfur [69, 70] and oxygen-based [156, 165-167, 198-

201] discharges.

Another example that deserves to be pointed out is the detailed study of Granier et al. dealing

PT
with the investigation of plasma chemistry by OES of hexamethyldisiloxane (HMDSO) and

RI
tetraethoxysilane (TEOS) discharges with and without oxygen [199]. It has been shown that

SC
HMDSO plasmas are dominated by Si, SiO and SiH species whereas OH, CO, CO+ and CO2+

emission lines are mainly identified in TEOS plasma. From these data, the authors concluded that

NU
the presence of CO and OH molecules in TEOS plasmas originates from surface reactions at the
MA
plasma/growing film interface followed by their desorption. These findings have allowed

providing relevant information on the growth mechanism of the coatings, an essential step for
D

understanding film properties. Nevertheless, it has to be stressed that given the complexity of
TE

organic plasmas in terms of the diversity of the molecules present in the discharge, a complete

identification of all species constituting the gas is extremely difficult by using solely OES.
P
CE

On the other hand, OES can also be used to monitor other important features of the plasma,

for instance the capacitive-to-inductive transition when working with inductively-coupled plasma
AC

discharges. This has been demonstrated in our group for the plasma polymerization of

propanethiol using a copper coil connected to an RF power supply as a plasma source [66, 132].

It is known that, in such a working environment, the capacitive discharge is characterized by a

weak global emission intensity since the energy transfer from the generator to the plasma

electrons is not efficient. On the other hand, as the plasma enters the inductive regime, i.e., by

increasing the RF power to the coil, the gas becomes much brighter. Fig. 11 shows how the

capacitive-to-inductive transition is accompanied by: i) an increase in the global emission

intensity, ii) an increase in the PPF deposition rate and iii) a decrease in the sulfur content in the

39
ACCEPTED MANUSCRIPT

PPF. Such behavior is related to the intense fragmentation of the precursor when more energy is

dissipated in the plasma.

To the contrary, it is difficult to obtain quantitative data through OES measurements, e.g.,

PT
determining the absolute density of species identified from the data displayed in Fig.10. The

RI
detected intensity of the line Ix, appearing at a wavelength λ, and emitted by an excited specie x

SC
can be expressed according to Eq. 5:

NU
I x = n x*Aijk(λ) , (5)

where nx* is the number density of the excited species x and k(λ) characterizes the optical setup
MA
response at the wavelength λ. Aij is the frequency of spontaneous emission of a photon at the

wavelength λ following the radiative decay from the the excited state j toward the lower energy
D
TE

state i.
P

In cold low-pressure plasmas, the production of excited levels typically involves electron
CE

collisions with the species. The rate constant of this collision depends on the collision cross

section σ(E), which is a function of the kinetic energy of the electrons, and of the EEDF. The
AC

later is therefore a key property that should be determined in order to extract quantitative

information from the OES measurements. Although one can find collision cross section data in

the literature, e.g., for elastic collisions with polyatomic molecules relevant to plasma processing

[202] and inelastic cross section for electron with hydrogen [203] or oxygen molecules [204],

measuring the EEDF is often problematic. Nevertheless, careful Langmuir probe measurements

can be carried out to provide this information (see previous section). Furthermore, if the atom or

molecule is excited to a metastable energy level, which is characterized by a much longer

radiative lifetime as compared to radiative excited states, another loss mechanism, involving

40
ACCEPTED MANUSCRIPT

diffusion outside the detection area (e.g., towards the chamber walls), must also be taken into

account. This situation would make a quantitative description of the plasma even more

complicated. This is the reason why OES should not be considered as the tool of choice to

PT
provide a quantitative insight on the plasma chemistry during plasma polymerization.

RI
SC
NU
MA
D
P TE
CE
AC

Fig. 11. Influence of the RF power on the capacitive-to-inductive transition for plasma

polymerization of propanethiol accompanied by: (a) an increase in the global emission intensity,

(b) an increase in the deposition rate and (c) a decrease in the sulfur content in the PPF. Inset:

Typical emission spectrum of the ICP discharge taken in the inductive mode for a power of 100

W [132].

Nevertheless, in some special experimental conditions, actinometry can be adapted to plasma

polymerization in order to estimate the relative concentration of species such as H, CH, CO, OH,

41
ACCEPTED MANUSCRIPT

and CN. Examples of such analyses can also be found in the literature for nitrogen [192, 205-

208] fluorine [197], hydrocarbon [168] and oxygen [156, 166, 167] containing discharges.

The actinometry method is explained in the book of Lieberman and Lichtenberg [92]. Briefly, the

PT
technique consists in calculating the line intensity ratios Ix/It of two plasma species. The first one,

RI
Ix, is related to the particle of unknown concentration, the other one, It, is emitted by the tracer (of

known concentration). The excitation threshold energy and the excitation cross section of these

SC
two emitters must be nearly the same. In this way, the ratio of intensities Ix/It is directly

NU
proportional to the concentration of the emitting species nx. This line intensity ratio is equal to:
MA
IX n X n e K X
= , (6)
It n t n e K t
D

In Eq. 6, ne is the electron density and Kx, Kt are related to:


TE

i) The production rate of the considered excited states (considering electron impact
P

excitation, the later depend on the EEDF and the electron-specie collision excitation
CE

cross section),
AC

ii) The radiative lifetime of the considered excited species, and

iii) The global response of the optical setup (e.g., the transmission of the lenses, optical

fibers, the sensitivity of the detector, etc. depend on the wavelength).

If the tracer t is properly chosen, the line ratio is proportional to nx. Since Kx, Kt, ne cancel

out and nt is known, nx can be deduced. In Ref. [168], actinometry was implemented to quantify

the production of H2, H and CH species in an Ar/styrene plasma. The fragmentation of the

aromatic precursor was found to increase with the RF power delivered in the plasma (Fig. 12). In

their publication, Choudhury et al. varied the RF power from 20 W to 130 W. They found that

42
ACCEPTED MANUSCRIPT

the properties of the styrene – based PPF are improved at a RF power of 100 W. OES and film

characterization data pointed out that improvement at this specific value of the power is due to

the predominance of CH radicals in the plasma and an enhanced cross-linking density due to the

PT
presence of a highest percentage of carbon content in the film (Fig. 12). These researchers

RI
concluded their study by suggesting the possibility of using styrene-based PPF, deposited at RF

power of 100 W, as high performance protective coatings for metal surfaces. Using a similar

SC
methodology, Palumbo et al. has highlighted an inverse correlation between the relative

NU
proportion of CO in the discharge and the retention degree of the precursor in acrylic acid PPF

[167].
MA
In another study, Bousquet et al. used the actinometry method for investigating the temporal
D

evolution of the relative density of atomic oxygen by time-resolved OES measurements in


TE

HMDSO/O2 plasmas [166]. In order to monitor the reactive species during the post-discharge

(i.e., when light is no longer emitted), the authors employed a short pulse excitation technique.
P
CE

Briefly, this method consists in applying a second shorter probing pulse at a RF power similar to

the one applied during the main pulse. The aim is to create electrons for exciting the remaining
AC

long-lived species. More details can be found in Refs. [166, 209]. This approach has allowed

measuring the O-atom loss coefficient on surfaces as a function of the chemical composition of

the gas. Furthermore, it has been shown that deposition events during the plasma “OFF” time

occur due to the dissociation of HMDSO molecules by long-lived oxygen atoms.

43
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
Fig. 12. Relative concentration of H, H2 and CH species in an Ar/styrene plasma, as a function of

the RF power injected to the plasma [168].


D
TE

Although extensively used in the field of organic plasma diagnostic, most of the time, the

OES technique does not allow drawing a complete picture of the plasma chemistry. Therefore, its
P
CE

impact in the understanding of the growth mechanism is limited to specific cases. However, the

relatively low cost of the equipment and its ease of implementation justify the use of the OES
AC

technique for a rapid screening of plasma chemistry.

44
ACCEPTED MANUSCRIPT

4.3 Mass Spectrometry

As explained in the previous section, only a partial picture of the chemical composition of the

plasma can be accessed from OES. In this context, the mass spectrometry (MS) technique has

PT
emerged as the most widely employed diagnostic method for investigating the plasma

RI
polymerization process. Indeed, in comparison to the OES method, a deeper knowledge of the

SC
chemical composition of the discharge including ions, stable molecules and radicals can be

obtained. The first reports of the analysis of organic plasmas by MS date from the 1980s for

NU
polymerizing/etching fluorine based-discharges [197, 210]. Nevertheless, the use of the mass
MA
spectrometry technique as a tool for a mechanistic understanding of the growth of functionalized

PPF has expanded in the 1990s with numerous contributions coming from the group of R.D.
D

Short in Sheffield [38, 57, 102, 169, 170].


TE

In order to avoid: (i) failure of some electronic part of the apparatus and (ii) collisions of
P

analyzed ions during their transport inside the equipment, a pressure less than 5.10-6 mTorr is
CE

necessary in the MS [175]. Therefore, for plasma polymerization processes operating at a

pressure less than 100 mTorr, the MS is connected to the deposition chamber through a small
AC

orifice (generally 100 µm in diameter) and is independently pumped while for plasma

polymerization conducted at higher pressure, a multistep differential pumping is required [211].

Depending on the analysis mode, this technique enables the detection of neutral (residual gas

analysis, RGA mode) and ionic (glow discharge mass spectrometry, GDMS mode) species

present in the gas phase. Since the mass analyzer only discriminates ions according to their mass-

to-charge ratio (m/z), the sampling of neutral species requires an ionization source. The most

common approach consists in heating a filament to generate an electron beam normal to the

45
ACCEPTED MANUSCRIPT

neutral particle flow; this results in ionization of the neutral species by electron impact and leads

to the formation of cations. Obviously, the kinetic energy of the incident electrons has to be

higher than the threshold ionization energy of the molecules (typically 10 eV for organic species).

PT
It should be noted that electron attachment processes can also be employed leading to the

RI
formation of negative ions [212]. In this case, the kinetic energy of the electrons is lower than the

threshold ionization energy. Nevertheless, this ionization method is only restricted to highly

SC
electronegative species (e.g., fluorine-based) and has therefore been significantly less employed.

NU
We will therefore focus only on the electron impact ionization mode for the detection of neutrals.
MA
For both analysis modes of the spectrometer (i.e., RGA and GDMS), the ions are separated

according to their m/z ratio by means of a quadrupole analyzer, the most commonly used mass
D

analyzer for plasma analysis [175]. Its working principle relies on the combination of DC and
TE

RF potentials applied to four conducting rods to define stable trajectories in oscillating electrical

fields which allow separating ions according to their m/z ratio. It should be noted that the
P
CE

quadrupole analyzer is characterized by a transmission function which strongly decreases with

mass. For most of mass spectrometers, the manufacturer found empirically that this function
AC

ranges between m-1 and m-2. In the literature, an m-1 correction factor is generally applied to the

mass spectra [41, 47, 101, 104, 128, 170]. Once the ions have been separated according to their

m/z ratio, they are converted to a measurable signal using a secondary electron multiplier (SEM)

detector. More details about the instrumentation can be found in the book of de Hoffmann and

Stroobant [213] and the review of Benedikt et al. [175]. For the GDMS mode, the technique also

enables time and energy-resolved measurements. In contrast, in RGA mode, time-resolved

measurements can not be obtained since the transit time of neutral species across the plasma-

46
ACCEPTED MANUSCRIPT

instrument boundary layer and the ionization chamber is typically longer than the pulse duration

when the process is operated in pulsed mode [41].

4.3.1 Neutral species analysis

PT
With regard to the RGA detection mode, a non-negligible drawback is related to the

RI
ionization process taking place in the ionization source of the equipment. Indeed, the ions are

SC
formed in an excited state and the release of excess of energy may lead to the fragmentation of

NU
the molecular ion, hence resulting in the appearance of additional peaks in the mass spectrum as

schematically described in Fig. 13. In other words, the detection of a signal in the mass spectrum
MA
does not necessarily imply that the corresponding species are formed in the plasma. In

conventional MS, a kinetic energy of 70 eV is usually employed in order to maximize the signal
D

intensity since for organic molecules, the maximum of the electron beam ionization cross section
TE

is located around 70 eV [175, 213]. Nevertheless, for plasma analysis, an electron kinetic energy
P

of 20 eV is preferable in order to limit excessive precursor fragmentation in the ionization source


CE

of the mass spectrometer [41, 47, 48, 54, 55, 66, 68, 131]. This value corresponds to the best

compromise between reduction of fragmentation in the ionization source and the threshold
AC

ionization energy of various organic-based molecules and fragments [41].

47
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
Fig. 13. Schematic description of the principle of mass spectrometry in RGA mode. The
MA
ionization process taking place in the ionization source induces, for some species, the

fragmentation of the parent ion, hence resulting in the appearance of additional peaks in the mass
D

spectrum (in red).


TE

Based on previous considerations, when operating in the RGA mode, the main challenge is
P

then to differentiate the molecules/fragments produced through dissociation/rearrangement


CE

reactions in the plasma from those formed in the ionization source of the spectrometer. Although
AC

this aspect is often neglected, a very simple relation (Eq. 7) can be applied to subtract the

fragmentation undergone by the precursor in the apparatus itself following [41, 54, 68, 131]:

IMonomer (Plasma ON)


Ic (m)  Im (Plasma ON) - Im (Plasma OFF). , (7)
IMonomer (Plasma OFF)

where Ic (m) is the calculated peak intensity for m/z = m, Im (Plasma ON) and Im (Plasma OFF)

represent the experimental peak intensity for mass m when the plasma is switched ON and OFF,

respectively. IMonomer is the intensity corresponding to the precursor signal.

48
ACCEPTED MANUSCRIPT

An illustrative example of the application of Eq. 7 is shown in Fig. 14 for the plasma

polymerization of acrylic acid. Together with allylamine, this PPF is probably the most studied in

the literature. Fig. 14 a-b present the mass spectra of acrylic acid (in the absence of plasma) and

PT
when the plasma is switched ON in pulsed mode (50W, ton = 100 µs, toff = 1000 µs), respectively.

RI
The peaks can be assigned as follows [38, 163]: m/z = 72 the acrylic acid precursor; m/z = 55 to

CH2CHCO+; m/z = 44 to C2H4O.+; C3H8.+ and CO2.+; m/z = 28 to C2H4.+ and CO.+; m/z =27 to

SC
C2H3+; m/z = 26 to C2H2.+; m/z = 18 to H2O.+; m/z = 2 to H2.+; and m/z = 1 to H+. The mass

NU
spectrum of the plasma corrected according to Eq. 7 is depicted in Fig. 14c. After the correction,

peaks previously identified at m/z = 55 and 27 almost disappear, indicating that these signals
MA
originate from dissociations reactions in the ionization source. This example illustrates the

importance of the correction method for obtaining the right picture of the neutral plasma
D

chemistry.
P TE
CE
AC

49
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 14. Neutral mass spectra of acrylic acid: (a) plasma OFF, (b) plasma ON (100 µs on-time,

1000 µs off-time, 50W), (c) corrected for both fragmentation of the acrylic acid precursor in the

ionization source according to Eq. 7 and mass transmission [41].

50
ACCEPTED MANUSCRIPT

In Fig. 14, it is worth noting the presence of numerous peaks in the mass spectrum indicating

a complex neutral-gas phase chemistry even at low power as also reported for other plasma

polymer families [47, 55, 66, 104, 170, 214]. Although traces of molecules with m/z higher than

PT
the precursor have been observed in some studies, no “dimer” or “trimer” based on the

RI
monomeric repeat unit is identified in contrast to positive ions (see hereafter) [37, 66, 104]. One

exception is the plasma polymerization of methyl isobutyrate where a peak associated with a

SC
neutral dimer has been observed [215].

NU
A closer look to the mass spectrum in Fig. 14 also reveals the production of stable
MA
hydrocarbon based molecules (e.g. C2H2, C2H4), as frequently encountered whatever the

precursor employed [38, 47, 54, 55, 104, 128]. As proposed by Thiry et al. and supported by
D

theoretical calculations, these hydrocarbon molecules result from rearrangement reactions of the
TE

radicals produced from dissociation reactions in the gas phase [66]. In addition to the formation

of stable hydrocarbon-based molecules, stable molecules containing a heteroelement (CO, CO2


P
CE

and H2O in the present example) are also formed [37, 38, 128]. Similarly, NH3 and N2 species are

identified in nitrogen-based discharges [47, 48, 104, 170] as well as H2S and CS2 in sulfur
AC

containing organic plasmas [66, 68-70, 216]. The concentration of such molecules in the gas

phase is generally inversely correlated to the heteroelement concentration in the films [217].

Indeed, considering that these unreactive molecules can not be trapped within the PPF, these

species do not take part in film growth, hence reducing the amount of reactive molecules

containing the heteroelement available for the deposit.

Another complication for the treatment of mass spectrometry data is related to the poor mass

resolution encountered for the quadrupole analyzer, i.e., 1 atomic mass unit. This leads to

numerous isobaric interferences (i.e., several species contributing to the same ion signal). For

51
ACCEPTED MANUSCRIPT

example, the peak at m/z = 28 in Fig. 14 can be attributed to C2H4 and/or CO. To overcome this

problem, several strategies can be employed. In their work, Haddow et al. used a plasma of both
13
C-labeled and unlabeled acrylic acid in order to facilitate the attribution of the peaks in the mass

PT
spectra [38]. In some particular cases, we can take advantage of the isotopic abundance of the

RI
heteroelement to discriminate molecules presenting similar m/z ratios, as shown in the case of

propanethiol discharges [66]. Finally, another strategy consists in measuring the intensity of the

SC
signal at a given m/z as a function of the kinetic energy of the colliding electrons [198]. This

NU
method enables one to deduce the ionization energy of the specie(s) contributing to a given signal

and then to determine their chemical composition. Ultimately, the combination of other diagnosis
MA
methods such as gas phase Fourier transform infrared spectroscopy and OES can also assist the

interpretation of the mass spectra data [53, 200].


D
TE

Although the intensity of a signal in an RGA mass spectrum is related to the density of the

corresponding species, calculation of their absolute concentration is not straightforward. Among


P
CE

other things, this requires the knowledge of the ionization cross-section which, for many stable

molecules, are available in the literature [175]. In this case, the absolute density of the probed
AC

molecules can be obtained by means of a calibration gas with a known density and electron

impact ionization cross-section [175]. With respect to radical species, determination of their

absolute density is much more demanding as they can be lost through collisions with the walls of

the reactor [173, 175]. Most of the time, the corresponding cross-sections are not available and

several assumptions have therefore to be made [175]. Furthermore, in some cases, specific

experimental configurations such as multistep differential pumping are needed for an accurate

measurement [173, 174]. Based on these considerations, in the context of the plasma

polymerization process, the great diversity of reactions (including fragmentations,

52
ACCEPTED MANUSCRIPT

rearrangements, etc.) highly complicates measurements of the absolute densities of the numerous

species constituting the plasma.

Nevertheless, some simple methods can be employed for obtaining quantitative or semi-

PT
quantitative information. For instance, the absolute precursor concentration can be easily

RI
measured using a calibration curve relying on the intensity of the monomer signal as a function of

SC
the pressure when the plasma is switched OFF [41]. Using this calibration curve, the intensity of

the precursor measured in the mass spectrometer, whatever the plasma parameters, can be

NU
directly related to the partial pressure of the monomer. For example, a decrease in the partial
MA
pressure of the acrylic acid precursor as a function the energy delivered in the plasma has been

reported [37, 38, 41]. This expected behavior is attributed to an increase in ne with power
D

resulting in an increase in precursor fragmentation [142]. It should be noted that the amount of
TE

unfragmented chemical precursors remaining in the discharge is generally related to the

concentration of the chemical group hosted by the monomer [47, 48].


P
CE

To monitor the relative proportion of radicals and stable molecules produced through

fragmentation/rearrangement reactions in the plasma, a depletion function can be employed


AC

according to Eq. 8 [172, 218]:

I -I   I 
D (%) =  OFF ON  x 100 = 1- ON  x 100 , (8)
 IOFF   IOFF 

where ION is the intensity of a peak at a given plasma condition and IOFF is the intensity of the

same peak in the absence of a plasma.

Based on Eq. 8, when a peak intensity increases with power, the depletion of that peak is

negative. Negative values indicate that the corresponding species are formed by the action of the

53
ACCEPTED MANUSCRIPT

plasma. In other words, the higher the negative value of the depletion function, the higher is the

production of the corresponding species in the plasma. Using this methodology, Hazrati et al.

have studied the formation of key fragments for the plasma polymerization of ethanol and

PT
demonstrated the production of radicals such as .CH3 and .OH to the detriment of the precursor

RI
molecule upon increasing the power [172]. It is important to stress here that Eq. 8 has to be

employed with care depending on the plasma polymer investigated. Indeed, species formed from

SC
the dissociation of the precursor ion in the source of the spectrometer affect the measured

NU
intensities and therefore the conclusions drawn from the depletion function.
MA
Another strategy has been recently implemented for directly correlating plasma and film

chemistries in view of the elucidation of the growth mechanism of propanethiol PPF [68].
D

Briefly, under certain experimental conditions, it has been shown that propanethiol PPF contains
TE

a sulfur content much higher than in the precursor. This particular feature is explained by the

trapping of H2S molecules within the plasma polymer matrix [66, 131, 132, 219]. To validate this
P
CE

hypothesis, the relative proportion of H2S (Irel. H2S) with respect to the sum of signals associated

with carbon-based species (m/z = 15,26-29,39-43), which could potentially take part to the
AC

growth of the PPF, has been calculated. The chosen species refer to radicals or molecules

containing an unsaturation which could potentially be grafted at dangling bonds present at the

growing film surface. Irel. (H2S) has then been calculated following Eq. 9 and compared to the

sulfur to the carbon ratio (S/C) measured by XPS in the corresponding PPF (Fig. 15):

IC (H 2S)
Irel. (H 2S) = m/z  29 m/z  43
, (9)
IC (m / z  15)  
m/z  26
IC (m / z)  
m/z 39
IC (m / z)

54
ACCEPTED MANUSCRIPT

where Ic (m/z) corresponds to the corrected intensity calculated following Eq. 7 for the signal at

m/z.

The obtained linear correlation between Irel. (H2S) and the S/C ratio as a function of power

PT
validates the trapping hypothesis, revealing the attractiveness of the mass spectrometry technique

RI
for a deeper understanding of the PPF growth at a molecular level.

SC
NU
MA
D
P TE
CE

Fig. 15. Evolution of the sulfur to carbon ratio of propanethiol PPF as a function of Irel. (H2S) for

different powers [68]. See the text for details.


AC

We understand from these examples that the interpretation of mass spectrometry data is

often challenging. As already mentioned, one strategy to make it easier consists in exploiting

DFT calculations. As an example, the study of the plasma polymerization of ethyl lactate, finding

applications in the design of biodegradable coatings, is presented [54]. Briefly, the objective of

this work was to correlate the ester content in the films, a key parameter for controlling the

biodegradability of the material, with the plasma chemistry. A particular feature of the mass

spectra of ethyl lactate plasma in RGA mode is the absence of a signal corresponding to the

55
ACCEPTED MANUSCRIPT

precursor (m/z = 118), thus making the interpretation of the data highly complicated. This is

explained by the “brittleness” of the ethyl lactate molecule which suffers of extended

fragmentation reactions in the ionization source of the spectrometer. Based on mass spectrometry

PT
data, a thorough mechanistic study using DFT was undertaken. A detailed description of the

RI
complex calculated fragmentation pathway occurring in the ionization source of the spectrometer

as well as in the plasma (Fig. 16 a-b) is beyond the scope of this review and interested readers

SC
can consult Ref. [54]. We will concentrate on the species at m/z = 75 (i.e., C3H7O2+) which

NU
deserve peculiar attention (see highlighted rectangle in Fig. 16). Using the DFT calculations, it

can be learned that this species can only be formed through an ionic mechanism involving the
MA
undamaged ethyl lactate ion and thus takes place in the ionization source of the spectrometer

(Fig. 16 a). Indeed, considering a radical mechanism occurring preferentially in the plasma (Fig.
D

16 b), the intermediate of the corresponding reaction is unstable and rearranges spontaneously to
TE

a more stable form at m/z = 74. Comparing both fragmentation patterns (i.e., ionic and radical),
P

this implies that the intensity of the peak at m/z = 75 is proportional to the density of undamaged
CE

precursor in the plasma. Therefore, this peak can be correlated with the evolution of the ester
AC

content in the PPF depending on the synthesis conditions. This example highlights the

significance of theoretical tools for extracting relevant information from complex diagnostic data.

A similar approach has been successfully employed for deriving the chemical reactions pathways

encountered by the precursor in propanethiol [66, 68] and benzene/cyclohexane [220] plasmas.

56
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 16. Proposed fragmentation pathways based on DFT calculations of ethyl lactate dissociation

in (a) the ionization source of the spectrometer and (b) in the plasma. Adapted from [54].

Another example revealing the strong interest in combining DFT calculations with MS data

is related to the plasma polymerization of allylamine and cyclopropylamine. The aim of this work

was to study the role of the nature of the precursor on the density of primary amines group in the

57
ACCEPTED MANUSCRIPT

PPF [48]. It has been shown that above a critical value of the mean power (around 30W)

delivered in the plasma, both RGA mass spectrometry and XPS measurements indicate that the

two precursors behave in a very similar way, showing significant fragmentations and poor

PT
incorporation of primary amines in the films. However, below this value, the content of amine

RI
groups increases in the films in a larger proportion for the cyclopropylamine precursor. These

data have been understood based on DFT calculations. Fig. 17 displays the enthalpies of

SC
reactions associated with various initial fragmentation schemes of allylamine and

NU
cyclopropylamine, as calculated by DFT. Note that: (i) all calculations have been performed in

the unrestricted scheme and, (ii) activation and bond dissociation energies are almost equivalent
MA
for bond dissociation processes [221]. In the case of allylamine, the more thermodynamically

favorable reaction involves the rupture of the C-N bond to form allyl and NH2 radicals. The lower
D

bond dissociation energy is rationalized by the stabilization of the allyl radical by resonance
TE

effects. In contrast, the easiest reaction in the case of cyclopropylamine is the opening of the ring,
P

thus retaining the amine chemical functionality on the precursor. This different behavior is fully
CE

consistent with the increased amount of amino group detected in the film upon plasma
AC

polymerization of cyclopropylamine. Accounting for the entropic effects would systematically

decrease by about 0.5 eV, the calculated bond dissociation energies while keeping unchanged the

previous conclusions. The full consistency between the experimental and theoretical results thus

paves the way toward a theoretical screening and even the design of precursors aimed at

optimization of the degree of PPF functionalization.

58
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE

Fig. 17. DFT-calculated bond dissociation energies of the initial fragmentations of allylamine
CE

(top) and cyclopropylamine (bottom). Adapted from [48].


AC

4.3.2 Ion analysis

As already mentioned, the ions also play a major role in the PPF growth mechanism. The

ionic gas phase chemistry has therefore also been extensively investigated by MS. Fig. 18

represents a typical example of the chemical composition of positive ions in an acrylic acid

discharge. Similar to the RGA analysis mode, numerous peaks are identified. It is important to

stress that in this case, the ions during their transport in the mass spectrometer obviously do not

suffer of fragmentation before their detection. Therefore, all peaks appearing in the spectrum

correspond to ions formed in the plasma. The base peak at m/z = 73 is ascribed to the protonated

59
ACCEPTED MANUSCRIPT

monomer ([M+H]+) while the most prominent fragments are observed at m/z = 39 (C3H3+), 55

(CH2CHCO+) and 57 (CH3CH2CO+). A relevant observation is the detection of oligomeric ions of

the form (2M+H)+ at m/z = 145 and (3M+H)+ at m/z = 217, where M corresponds to the

PT
molecular weight of the starting material as also observed in other works [101, 128, 215, 222].

Oligomer ions corresponding to [4M+H]+ have also been identified in Ref. [38]. The dimer and

RI
trimer ions can lose H2O giving rise to additional signals separated by a value of 18. Such

SC
oligomer ions formed through gas phase neutral/ion reactions have also been identified in

NU
propanoic acid [215, 222], allylalcohol [214], propanol [214], allylamine [104, 170], HMDSO

[102, 223, 224], methyl isobutyrate, methyl methacrylate, n-butyl methacrylate [169], ethanol
MA
[172, 225], γ-terpinene [171] and thiophene [218] plasmas. Most of the time, the formation of

these oligomeric ions involves the addition of a hydrogen ion on the neutral precursor followed
D

by the successive additions of uncharged monomers. However, the exact mechanism remains
TE

unclear in some aspects. For example, gas phase oligomerization reactions do not require the
P

presence of a double bond in the organic precursor marking a clear difference with classical ionic
CE

polymerization [169, 215]. The reason why neutral/radical addition reactions in the gas phase is
AC

ruled out lies in the kinetics of the reactions as proposed by Benedikt [226]. Indeed, ion-neutral

reactions are typically 10 times faster than reactions between neutral particles due to the

attractive potential created by the induced dipole moment in the neutral particle.

60
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 18. Positive-ion mass spectrum of acrylic acid plasma sustained at a power of 3W. The

region under the horizontal bar has been expanded by a factor of 5 in intensity. Adapted from

NU
[215].
MA
The intensities of the peaks associated with the oligomeric cations were found to evolve with

the power dissipated in the discharge. Most of the time, a decrease in the relative proportion of
D

these ions in the plasma upon increasing the power has been reported [37, 169, 171]. This
TE

behavior can be ascribed to a decrease in the concentration of the precursor, hence reducing the
P

probability of addition reactions with ions. Since a correlation between the relative proportion of
CE

ions and the degree of retention of the precursor in the films has been highlighted for several PPF
AC

systems, it has been proposed that these ions significantly contribute to the film formation

especially at low power [128]. It should be noted that for allylamine plasmas, a nearly constant

proportion of oligomeric ions has been reported whatever the power. This points out the influence

of the chemical nature of the precursor on the oligomerization reactions [104, 170].

For sake of completeness, negative ions (fragments and oligomers) have also been identified

in silane [227] and oxygen containing discharges [228, 229] using MS. Their detection is only

possible in the afterglow at sufficient long OFF time for enabling their extraction from the bulk

plasma. In continuous wave (CW) plasma polymerization, their contribution in the condensing

61
ACCEPTED MANUSCRIPT

material is excluded since they are confined in the bulk of the gas due to the positive electrical

potential drop between the plasma and the surrounding surfaces. When operating the discharge in

the pulsed mode, they are able to reach the film surface only for sufficient long plasma “OFF”

PT
times. Their role in the overall deposited mass has been estimated to be quite low (< 1%) in

RI
comparison to positive ions and neutrals [228].

SC
Although the pulsed mode has been extensively employed in the plasma polymerization

domain, only a few studies were dedicated to time-resolved mass spectrometry measurements. As

NU
aforementioned, such kinds of measurements are limited to ionic species [41]. The results
MA
obtained have allowed a better understanding of the chemistry of pulsed organic discharges. The

main conclusions drawn are that some ions can “survive” during the plasma “OFF” time during at
D

least 1000 µs after the extinction of the plasma [41, 163]. This aspect is illustrated in Fig. 19
TE

where GDMS measurements were performed in acrylic acid discharges for selected times during

the plasma “OFF” time. The dominant positive ions are identified as follows: m/z = 55 to
P
CE

CH2CHCO+, m/z = 73 to [M+H]+ (M being the acrylic acid precursor), m/z = 127 to [M+H-

H2O]+, m/z = 145 to [2M+H]+, m/z = 199 to [3M+H-H2O]+ and m/z = 217 to [3M+H]+. After 500
AC

µs, only the trimeric species (at m/z = 217) are still detected in the plasma and remain observable

after 1000 µs. In addition, one can also notice that the intensity of the ions at m/z = 199 and 217

are higher after 500 µs than 200 µs. This clearly indicates the possibility of OFF-time reactions of

monomers with dimers to form trimers within the time scale of the OFF-period. These results

have challenged the usual view of a plasma polymerizing pulsed discharge, namely a

concomitant ionic and neutral deposition mechanism during the plasma ON time and an

exclusively neutral chemistry during the OFF time.

62
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 19. Time-resolved ion mass spectra of acrylic acid discharges obtained at 50 W RF power

(ton = 100 µs and toff = 10000 µs). Adapted from [41].

63
ACCEPTED MANUSCRIPT

In addition to the significance of determining the neutral/ionic chemical composition of the

plasma, investigating the energy of the ions bombarding the substrate is also of considerable

interest for a mechanistic description of plasma polymer formation. Indeed, by means of a

PT
specific RF biasing technique enabling to control the ion energy independently of other

RI
parameters, Barton et al. have reported that the ion energy can significantly alter some important

film properties such as the deposition rate as well as the layers chemistry [222, 230]. The mass

SC
spectrometry technique allows us to measure the energy distribution function of the ions (IEDF)

NU
arriving at the orifice of the instrument which can be grounded or at floating potential. However,

it has to be mentioned that energy measurement at a floating orifice is much more representative
MA
of the energy of ions bombarding the growing film when the substrate is not connected to an RF

power supply since most of the time, plasma polymers are insulating [38]. For an accurate
D

estimation of the bombarding ion energy, the orifice of the mass spectrometer has to be localized
TE

at the position of the substrate. Fig. 20 represents a typical example of an IEDF recorded in a γ-
P

terpinene plasma using a spectrometer with a grounded orifice. The IEDF has a shape which is
CE

typical of those recorded in plasmas characterized by a collisionless sheath. The “peak” in the
AC

IEDF (at about 16 eV in Fig. 20) corresponds to ions accelerated across the plasma pre-sheath

and sheath and entering the MS without any collision (and hence without loss of energy) [171,

225].

64
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
Fig. 20. Energy distribution function at the grounded orifice of the mass spectrometer of the

precursor in γ-terpinene plasma sustained at a power of 50 W [171].


D

Haddow et al. have performed a systematic study of the evolution of the ion energy as a
TE

function of the dissipated power in CW plasma polymerization of acrylic acid using an external
P

RF inductive coil as plasma source [38]. The results are summarized in Fig. 21 depicting an
CE

increase in the ion energy from 5 eV to 30 eV when the power evolves from 1 to 15 W. At 3 W,
AC

the number of atoms in the molecular ion has been estimated to be, on average, 10 whereas the

ion energy is 10 eV, thus corresponding to an energy per atom <1.3eV [38, 101]. It is reasonable

to assume that at this value, fragmentation of the striking ion (and surface) will be minimal

enabling a better preservation of the function of interest. At higher power, the peak ion energy is

30 eV while the ions contain <10 atoms on average, leading to an energy per atom > 3 eV. Since

bond energies in organic molecules typically range from 3-5eV, the fragmentation of the

incoming ion and/or the growing film surface might be anticipated at higher powers, therefore

contributing to the decrease in the degree of retention of the precursor in the coating [38, 104].

65
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 21. Evolution of the ion energy of acrylic acid cations relative to a self-biased, or floating,

NU
surface as function of the power. Adapted from [38].
MA
To summarize, the popularity gained in the field of plasma polymerization by the MS

technique over several decades lies in its ability to determine relevant properties including the
D

chemical composition of the plasma and the energy of bombarding ions. This is probably the
TE

reason why this method has emerged as one of the most powerful diagnostic methods for

deepening the fundamental understanding of the mechanistic formation of a given PPF. The main
P
CE

disadvantage of the MS technique is related to the cost of the equipment and the often difficult

interpretation of the data. In the latter, it is clear that theoretical support such as DFT calculations
AC

can help.

4.4. Ion Probes

Using the MS technique, although all ions constituting the plasma can be identified, the

absolute ion flow toward the growing film cannot be deduced since MS measurements do not

give access to the absolute density of ions. This aspect is, however, of crucial importance to

determine the mass delivery by ions to the film surface and also in combination with other

techniques to measure the contribution of ions in the total energy flux reaching the growing film.

66
ACCEPTED MANUSCRIPT

In plasma processing, a classical means to measure the ion flux to a surface is based on ordinary

electrostatic probes inserted into the plasma volume. From measurements of the ion density and

electron temperature in the plasma bulk, the ion flux toward the interface is deduced using the

PT
Bohm criterion. Nevertheless, in the context of plasma polymerization, this method is not well

RI
suited for reliable and routine analysis. Indeed, as explained in section 4.1, the coverage of the

probe with an insulating PPF highly complicates the measurements [8, 101, 104, 176].

SC
In 1996, Braithwaite et al. have developed a novel probe to measure the absolute ion flux

NU
toward a surface in organic discharges which presents significant advantages: tolerant to
MA
insulating deposits, non-perturbing, easy to implement and applicable whatever the plasma source

[176]. Briefly, the basic principle consists in applying a pulsed RF voltage to the electrode.
D

Owing to the different mobility of electrons and ions, the probe naturally acquires a negative self-
TE

bias potential when an RF signal is applied. The RF signal is then chopped for approximately 5-

10 ms and the change in the voltage is measured with time as the ions impact the negatively
P
CE

biased electrode [8, 102]. The ion flux is deduced from the variation of the bias voltage as a

function of time. For more details about the physical phenomena involved in the measurement of
AC

the ion flux using this probe, the readers are invited to consult the paper of Braithwate et al.

[176]. Today, ion flux probes are relatively inexpensive and easy to use [58].

An example of the evolution of the ion flux toward the substrate as a function of the power is

displayed in Fig. 22 for the plasma polymerization of allylamine. The ion flux was found to

increase from 6.6x1016 m-2 s-1 ions at 1 W to 1.4x1018 ions m-2 s-1 at 14 W. Similar trends and ion

flux values were reported for acrylic acid [58, 222], propionic acid [58, 222], hydrocarbons (n-

hexane and 1,7-octadiene), diethylene glycol, diethylene glycol divinyl ether [58], γ-terpinene

[171] and ethanol [225] plasmas. The observed trend in Fig. 22 is ascribed to the increase in the

67
ACCEPTED MANUSCRIPT

electron density and, in turn, in the ion density in the plasma volume upon increasing the power

[142]. The calculation of the average ion mass based on the ionic mass spectra combined with ion

flux measurements allow us to deduce the ion mass flux toward the growing film. To evaluate the

PT
contribution of ions in the deposited mass, the latter is compared with the mass deposition rate

RI
(expressed as the total mass deposited per unit surface and time) obtained from quartz crystal

microbalance. The data are summarized in Table 2 for the present example. At low power (i.e.,

SC
1W), approximately 63% of ions could, in principle, contribute to film formation. Interestingly, at

NU
a power of 5W, the ion flux is sufficient to account for all deposited mass. This provides a strong

argument that ions significantly contribute to film formation. It is important here to stress that it
MA
does not mean that the film grows through a pure ionic mechanism and that neutral/radical

surface reactions are ruled out. Estimation of the mass delivery to the film by ions by a
D

comparison between the mass deposition rate and ion flux inherently takes into account that all
TE

ionic species exhibit a sticking coefficient equal to unity, or at least does not evolve with the
P

process parameters. However, the real situation is much more complicated as the sticking
CE

coefficient of ions is less than unity and is affected by their chemical nature, energy and the
AC

nature of the surface. For example, it can be learned from Table 2 that the ion flux exceeds the

mass deposition rate for a power of 5W, indicating a loss of ionic mass incorporated in the film.

A possible loss mechanism could be recombination events occurring at the growing film surface

followed by the formation of a stable molecule unable to be chemisorbed (e.g., CO.+ + e-surface 

CO (g) or CH3CH2CO+ + e-surface  CH3CH2 (surface) + CO (g)) [101]. It should be noted that a

more pronounced ablation phenomenon at higher power might also be anticipated. The previous

considerations also prevail for the sticking coefficient of neutral species for which it has been

reported that their surface reactivity is directly related to the density of dangling bonds at the

interface [99, 107]. Actually, both neutrals and ions are intertwined in the overall mechanism
68
ACCEPTED MANUSCRIPT

since an increase in the ion flux causes the formation of more radical sites at the surface, thus

resulting in a more efficient grafting of reactive neutral species.

PT
RI
SC
NU
MA
Fig. 22. Evolution of the ion flux in allylamine plasmas as a function of RF power [104].
D
P TE
CE
AC

69
ACCEPTED MANUSCRIPT

Power Total material deposition Positive–ion mass


rate flux
(W)
(µg m-2 s-1) (µg m-2 s-1)

PT
1 18.7 11.8

RI
3 61.7 36.2

SC
NU
5 86 99.4
MA
15 127.1 226.6
D

Table 2. Comparison of total mass deposition rates and positive-ion mass flux in allylamine
TE

plasmas for different powers [104]. See text for details.


P

The emergence of ion probes has also highlighted the influence of the chemical nature of the
CE

precursor on the process. Michelmore et al. investigated several saturated/unsaturated monomers,


AC

comparing the mass deposition rate with the ion flux as a function of power [57, 58]. For each

pair of precursors, despite similar ion fluxes for a given set of plasma parameters, the mass

deposition rate is significantly higher for the unsaturated monomer. This behavior is attributed to

a more pronounced ionic deposition mechanism for the saturated precursor, whereas the grafting

of intact precursor at open bond sites present at the interface is predominant for the unsaturated

one.

The development of the ion probe has therefore allowed shedding light on the contribution of

the ions to the PPF formation. Even though the exact role played by radicals and ions in the

70
ACCEPTED MANUSCRIPT

plasma polymer formation is still debated, it is however well-established that the actual picture of

the PPF growth mechanism should consider both species.

PT
RI
SC
NU
MA
D
P TE
CE
AC

71
ACCEPTED MANUSCRIPT

4.5 Gas-phase Fourier Transform Spectroscopy

As already mentioned, the mass spectrometry technique, is one of the most exploited

plasma diagnostic methods for probing PECVD discharges and, specifically, the plasma

PT
polymerization process [41, 48, 68, 172]. Nevertheless, it is accepted that it presents two major

RI
drawbacks, namely the fact that: (i) the quantification of species is complex such that most of

SC
time only semi-quantitative data can be obtained and (ii) the additional fragmentation of the

plasma generated species which likely occurs in the ionization chamber when using the RGA

NU
mode makes interpretation of the data even more complicated [99]. These limitations are in fact
MA
related to the “ex situ” feature of this plasma diagnostic technique. Therefore, in situ approaches

have been developed in order to overcome these drawbacks while taking care to avoid perturbing
D

the plasma during measurements.


TE

Optical techniques fulfil these requirements, especially the absorption spectroscopy (AS)
P

method because they allow determining population densities in both ground and metastable states
CE

as well as information related to the gas temperature when considering the line profile of atomic
AC

bands or the ro-vibrational structure of molecular gases [231]. The fact that absolute densities can

be deduced from AS measurements without instrument calibration is one of the main advantage

of this technique in comparison to OES, which was discussed in part 4.2. Among the different

wavelength regions that have been scanned, the infrared ( = 14 µm – 25 µm corresponding to

700-4000 cm-1) is particularly well adapted to plasma polymerization process because this region

of the electromagnetic spectrum contains the vibrational signatures of plasma-generated

molecular species.

72
ACCEPTED MANUSCRIPT

Historically, dispersive instruments were first used for IR absorption measurements, but

they are not well suited for detecting weakly absorbing molecular species in low pressure

plasmas. Therefore, today, plasma diagnostics using infrared absorption spectroscopy is often

PT
performed through Fourier transform infrared spectroscopy (FTIR) or using tuneable diode laser

RI
absorption spectroscopy (TDLAS). Both methods have their advantages and drawbacks. Laser

techniques typically offer higher sensitivity and resolution than FTIR, but cannot (even with

SC
tunable lasers) probe the broad bandwidth range required to track many IR absorption peaks

NU
simultaneously. An excellent review on the TDLAS spectroscopy can be found in Ref. [232].
MA
Initially, before FTIR and TDLAS have emerged as plasma diagnostic tools, IR

absorption was used to probe silicon [233] and a silicon dioxide dry etching processes [234],
D

using fluorocarbon gases to take advantage of the strong absorbing nature of the fluorocarbon
TE

species and thus to overcome the limited sensitivity of dispersive instruments [231]. With the

development of the above mentioned spectroscopic methods to probe low pressure plasma with
P
CE

sufficient sensitivity, infrared analysis of plasmas containing IR-sensitive species has become of

importance in view of the quantitative character of the analysis. Indeed, in molecular and low-
AC

temperature plasmas characteristic of the plasma polymerization process, the plasma-surface

interaction that governs thin film growth is controlled by the fragmentation of precursor

molecules in the plasma. Therefore, a better understanding of thin film growth strongly depends

on better evaluation of the plasma chemistry and reaction kinetics. This is only possible if we are

able to quantitatively probe the densities of species in the plasma. From an instrument point of

view, the sensitivity problem that is encountered in low-pressure discharges is often tackled by

implementing White-cell multiple pass optical arrangements. Fig. 23 illustrates, for example,

typical experimental setups used for FTIR (Fig. 23a) and TDLAS (Fig. 23b) measurements.

73
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE
AC

Fig. 23. (a) Experimental setup used (a) by Raynaud et al. to probe a microwave HMDSO plasma

by FTIR diagnostics [178] and (b) by Takahashi et al. to probe the discharge by TDLAS during

the plasma polymerization of fluorocarbon-based precursors [235].

Numerous PECVD processes involving hydrocarbons precursors to deposit C-containing

films such as Diamond Like Carbon layers or plasma polymers have been studied by IR

absorption spectroscopy in order to monitor transient and stable species that are strongly involved

74
ACCEPTED MANUSCRIPT

in the film growth mechanism. For example, Takahashi et al. have reported on the use of infrared

laser absorption spectroscopy to study the impact of the precursor (i.e., C4F8, C3F6, C5F8) on the

growth of fluorocarbon thin films by evaluating radical kinetics [235].

PT
RI
SC
NU
MA
D

Fig. 24. Power dependence of (a) CF, (b) CF2 and (c) CF3 radical densities measured by TDLAS
TE

for different fluorocarbon-based precursors [235].


P
CE

The coatings were synthesized in a capacitively-coupled plasma using pressures ranging

from 15 to 45 mTorr. In this work, they showed that the density of small mass radicals such as
AC

CF or CF2 (1010 and 1012 cm-3, respectively) are not impacted by the precursor nature in their

experimental conditions. However, stable molecules such as CF4 and C2F6 were generated from

C4F8 and C3F6 precursors suggesting a lower gas and surface polymerization for these precursors.

This is consistent with the higher growth rate and fluorine content obtained when using the C5F8

precursors [236]. Fig. 24 shows the power dependence of (a) CF, (b) CF2 and (c) CF3 radical

densities measured for the different fluorocarbon-based plasmas. These data, and particularly the

saturation of the CF2 signal with increasing power and the low densities of CF and CF3 radicals,

have allowed the authors to conclude that the contribution of these radicals to the film formation

75
ACCEPTED MANUSCRIPT

is not significant. Later, the same group extended this work by studying the co-polymerization of

C6F6 and C5F8 by OES and FTIR plasma diagnostics. In this work, they demonstrated that the

polymerization of C6F6 and its incorporation with a ring structure in the thin film improves the

PT
thermal stability of the films in comparison with pure C5F8 films [236].

RI
The PECVD deposition of C-containing coatings with other heteroelements than fluorine

SC
have also been studied. For example, Goujon et al., have studied plasma chemistry during

synthesis of SiOx coatings in a capacitively-coupled plasma using HDMSO/O2 mixtures at

NU
relatively high pressure (i.e., 1 Torr) by using OES and FTIR (without the use of a White cell)
MA
[200]. Their results have revealed, in agreement with previous work, the existence of two regimes

as a function of the HDMSO content in the gas mixture. They were also able to evaluate the
D

degree of fragmentation of the HDMSO precursor as a function of the injected power, which was
TE

a valuable input to proposing a tentative fragmentation pattern of the precursor. Similarly,

Raynaud et al. have studied the plasma polymerization of HMDSO in a microwave discharge by
P
CE

FTIR absorption spectroscopy [178]. In this case, the plasma was generated at low pressure (3

mTorr) using a Withe cell with up to 44 m of optical length (see Fig. 23a). These data were
AC

correlated to the chemical structure and composition of the deposited films in order to understand

the growth mechanism. It has been demonstrated that, surprisingly, the coating structure evolves

with an opposite trend in comparison to the plasma chemistry. At low power (Fig. 25a), the

growth mechanism is affected by the surface formation of stable volatile molecules (CH4,

(CH3)3SiH and pentamethyldisiloxane) from a significant part of the plasma generated radicals.

This implies that only a few basic radicals are responsible for the growth of films, especially

(CH3)xSiO. At high power (Fig. 25b), the plasma is dense enough to dissociate the by-products,

and hence to increase the flux of condensing species, leading to an increase in the Si-O grafting

76
ACCEPTED MANUSCRIPT

rate in the films. Although a significant insight into the growth mechanism has been obtained

thanks to plasma diagnostics using infrared spectroscopy, the authors mentioned that even if the

technique is powerful, the complexity of the spectra presenting numerous bands related to the

PT
different chemical species/radicals in the plasma makes interpretation difficult.

RI
SC
NU
MA
D
P TE
CE
AC

Fig 25. Simplified reaction pathway for (a) low and (b) high power conditions established from

FTIR diagnostic measurements for the plasma polymerization of HMDSO. Adapted from [178].

This later claim becomes even more relevant when considering the more complex plasma

polymerization systems that are today studied to synthesize functionalized organic surfaces. As

77
ACCEPTED MANUSCRIPT

an example, Wells et al. have studied the plasma polymerization of ethylene glycol in a pulsed

inductively-coupled plasma reactor by FTIR and OES in order to evaluate the dynamics of

monomer fragmentation and the effective chemical feedback from the boundary walls [59]. A

PT
pressure of 60 mTorr was used and the optical length was 64 cm. The presence in the plasma of

RI
CH, CO, OH and H (by OES) and of acetylene, ethylene, methane, water, formaldehyde, CO and

CO2 (by FTIR) was demonstrated. These stable volatile molecules are claimed to be the results of

SC
radical recombination and polymerization processes occurring at plasma/walls interfaces. It is

NU
also shown that evolution of the film chemistry as a function of the applied power is correlated

with the fragmentation pathway of ethylene glycol in pulsed plasmas deduced with a higher
MA
retention of the monomer functionality at low mean power.
D

The complexity of plasma polymerization processes clearly makes infrared data


TE

interpretation more and more difficult. In order to contribute to a better interpretation of these

data, our group has recently used a DFT strategy similar to the one mentioned earlier in the mass
P
CE

spectrometry section [53]. This approach has been developed in the context of the synthesis of

ethyl lactate plasma polymers. In order to validate our approach, we have first compared the
AC

infrared spectrum simulated for ethyl lactate with a corresponding experimental measurement

obtained from ethyl lactate vapor (plasma OFF), see Fig. 26. A comparison between the two

spectra show a very nice agreement, with a root-mean-square deviation of ~ 5 cm-1, hence

validating our theoretical approach.

When initiating the plasma, the fragmentation of precursor molecules generates a mixture

of different species leading to a much more complex spectrum. Mass spectrometry measurements

in RGA mode have allowed identifying nine dominant fragments: CH3, CH4, H2O, CO, CO2,

C2H4, C2H2, CH3CHOH, and C3H7O2. Fig. 27b exhibits the sum of the DFT-calculated infrared

78
ACCEPTED MANUSCRIPT

spectra of these nine fragments, with their relative intensities weighted on the basis of the peak

intensities from the mass spectrum. The weighting factor (WFi) for a given specie i is expressed

as:

PT
IiMS
WFi = (10)
 IiMS

RI
i

SC
with IiMS the intensity in the mass spectrum of the species i.

NU
This yields a simulated spectrum in marked discrepancy with the experimental one (Fig. 27a)
MA
due fragmentation in the ionization source of the spectrometer. The ester-bearing

molecules/fragments are not identified in the mass spectra and thus not considered to build the
D

synthetic IR spectrum. This can be corrected by estimating the percentage x EL of undamaged


TE

ethyl lactate molecules in the plasma (XEL) by comparing the intensity of the peaks related to the

ester function in the ethyl lactate vapor when the plasma is switched OFF and when a power of
P
CE

60 W is applied. This leads to a XEL value equal to 0.712. As a result, the peaks associated to the

fragments are further corrected by a factor (1- XEL) = 0.288. The resulting corrected spectrum,
AC

shown in Fig. 27c, appears to be in much better agreement with the recorded experimental

spectrum, allowing for the assignment of most of the peaks even though the presence of

fragments containing ester groups in the plasma has been neglected. This original protocol clearly

provides a new diagnostic tool to probe the plasma chemistry, especially the presence of specific

chemical functionalities.

In the framework of this research, we were also interested in the evolution of the ester-

containing species density as a function of the power which, in fine, partly determine the

degradation behavior of the deposited films. These data were used to better understand the

79
ACCEPTED MANUSCRIPT

growth mechanism of the ethyl lactate plasma polymer and, more specifically, to evaluate the

reaction efficiency of the ester-containing fragments (ester ) in the growing film [99]. The latter

was obtained by comparing the plasma chemistry to the film composition. Fig. 28 shows the

PT
evolution of ester as a function of the applied power. The measured trend highlights the higher

RI
incorporation efficiency of ester-based fragments in the growing film at high powers. This is

SC
likely related to a higher density of surface active site generated by ion and photon irradiation

upon increasing the power.

NU
MA
D
P TE
CE
AC

80
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE

Fig. 26. (a) Experimental IR spectrum of ethyl lactate in gas phase without discharge at 10 mTorr
AC

and 5 sccm and (b) IR spectrum of ethyl lactate calculated by DFT [53].

81
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
D
P TE
CE

Fig. 27. (a) Experimental IR spectrum of an ethyl lactate plasma sustained by a power of 60 W.
AC

(b) Synthetic IR spectrum calculated on the basis of Eq. 10 and (c) corrected according to the

degree of fragmentation of the precursor [53].

82
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 28. Evolution of the surface reaction coefficient of ester-bearing fragments (ester) as a

NU
function of power during the synthesis of ethyl lactate plasma polymers, as estimated from FTIR
MA
plasma diagnostic data. Adapted from [99].

Gas-phase FTIR spectroscopy is a powerful tool allowing for quantitative determination


D

of important species (radical, stable molecules) in the plasma phase. In combination with other
TE

diagnostic tools, as well as with theoretical calculations and surface analysis, it is even possible to
P

evaluate surface reaction coefficients. One of the important drawbacks is the complexity of the
CE

instrumentation which requires the design of dedicated reactors.


AC

4. Conclusions and perspectives

The interest in organic surfaces exhibiting a well-defined and tuneable chemistry with

stability is becoming increasingly important in many modern fields of application, especially in

the bio-technology sector. In this context, plasma polymer thin films historically used as

protection or barrier coatings have found a potential “New Age”. Nevertheless, these new

opportunities will only materialize if the necessary control of the physico-chemical properties of

the materials is attained. In view of the complexity of the plasma polymerization process, it is

todays accepted that such a control will only be possible through a deeper understanding of the

83
ACCEPTED MANUSCRIPT

plasma chemistry and ultimately of the plasma-surface interaction in order to determine

consistent growth mechanisms.

In this review, it is suggested that the required detailed analysis and understanding of plasma

PT
chemistry employed in low-pressure plasma polymerization processes can only be obtained by a

RI
complementary approach combining several state-of-the-art plasma diagnostic tools. The most

SC
employed tools are mass spectrometry (in both ion and neutral modes), optical techniques such as

OES and gas phase FTIR, and to a lesser extent electrostatic and ionic probes. Each of these

NU
techniques provides a piece of a puzzle that has to be handled with care. Indeed, in addition to the
MA
already challenging experimental measurements, the interpretation of the gathered data is, at

least, as difficult. Therefore, in order to access the physical values that will ultimately be used to
D

determine a convincing growth mechanism (e.g., flux of radicals, ions, sticking coefficients, etc.),
TE

the use of theoretical calculations such as those based on DFT can provide valuable support.
P

It is clear that the research community has understood the strong need for detailed knowledge
CE

of plasma chemistry as revealed by the numerous publications on this topic. Nevertheless, effort

is still required to go further in the direction of a quantitative description of this complex


AC

medium. In our opinion, effort is especially needed to better evaluate the surface reactivity of

neutral and ionic film-forming species with the growing film. The knowledge of this essential

physical parameter will ultimately makes possible the development of a new generation of

models allowing for the design of experiments in view of a given coating application. The wise

choice and even the design of precursors that can be driven by diagnostic tools and theoretical

calculation as demonstrated in some work reported in this review are a good examples of the

future knowledge-driven development of the plasma polymerization process.

84
ACCEPTED MANUSCRIPT

The future will also certainly see the confirmation of the rise of atmospheric plasma

polymerization processes. We intentionally did not consider these processes in the present

review, but we are clearly aware that they potentially represent the next step of development of

PT
this technology. Therefore, it is clear that diagnostic approaches, specific to the atmospheric

pressure plasma polymerization processes, should also be on the “to-do” list of the scientists

RI
active in the field.

SC
Altogether, we are convinced that large avenues are open for the use of plasma

NU
polymerization processes, but the price to pay will be more and more precise understanding of
MA
fundamental mechanisms which will only be obtained by a thorough evaluation of the plasma

chemistry using a combination of diagnostic tools.


D
TE

Acknowledgements
P
CE

D. Thiry thanks the “Région Pays de La Loire (France)” through the “Post-doctorats

internationaux” program and the FRIA grant from the “Fédération Wallonie-Bruxelles”
AC

for financial support program. S. Konstantinidis and J. Cornil are research associate and

director of research of the Fond National de la Recherche Scientifique (FNRS) Belgium,

respectively. D. Thiry, S. Konstantinidis and R. Snyders acknowledge the support of the

Belgian Government (Belspo) through the “Pôle d’attraction interuniversitaire” (PAI,

P07/14, “Plasma-Surface Interaction”, ψ) and of Wallonia Region (DG06) through the

program of excellence “OPTI2MAT” and the CONVERGENCE project “EVERWALL”.

85
ACCEPTED MANUSCRIPT

References

[1] K. Bewilogua, G. Bräuer, A. Dietz, J. Gäbler, G. Goch, B. Karpuschewski, B. Szyszka,

Surface technology for automotive engineering, CIRP Annals - Manufacturing Technology, 58

PT
(2009) 608-627.

RI
[2] C.D. Dimitrakopoulos, D.J. Mascaro, Organic thin-film transistors: A review of recent

SC
advances, IBM J RES DEV, 45 (2001) 11-27.

[3] M. Ohring, Materials Science of Thin Films, Elsevier Science, USA, 2001.

NU
[4] H. Biederman, Y. Osada, Plasma Technology, Elsevier Science Publishers, Amsterdam, 1992.
MA
[5] A. Hemberg, S. Konstantinidis, F. Renaux, J.P. Dauchot, R. Snyders, Ion flux-film structure

relationship during magnetron sputtering of WO3, The European Physical Journal Applied
D

Physics, 56 (2011) 24016.


TE

[6] K. Sarakinos, J. Alami, S. Konstantinidis, High power pulsed magnetron sputtering: A review

on scientific and engineering state of the art, Surf. Coat. Technol., 204 (2010) 1661-1684.
P
CE

[7] P.J. Kelly, R.D. Arnell, Magnetron sputtering: a review of recent developments and

applications, Vacuum, 56 (2000) 159-172.


AC

[8] A. Michelmore, D.A. Steele, J.D. Whittle, J.W. Bradley, R.D. Short, Nanoscale deposition of

chemically functionalised films via plasma polymerisation, RSC Adv., 3 (2013) 13540-13557.

[9] N.J. Saikia, C. Ewels, J.-F. Colomer, B. Aleman, M. Amati, L. Gregoratti, A. Hemberg, D.

Thiry, R. Snyders, C. Bittencourt, Plasma Fluorination of Vertically Aligned Carbon Nanotubes,

J. Phys. Chem. C, 117 (2013) 14635-14641.

[10] P. Chabert, N. Braithwaite, Physics of Radio-Frequency Plasmas, Cambridge University

Press, New York, 2011.

86
ACCEPTED MANUSCRIPT

[11] A. Grill, Cold Plasma Materials Fabrication: From Fundamentals to Applications, Wiley,

New York, 1994.

[12] K. Vasilev, J. Cook, H.J. Griesser, Antibacterial surfaces for biomedical devices, Expert

PT
Review of Medical Devices, 6 (2009) 553-567.

RI
[13] B.T. Houseman, E.S. Gawalt, M. Mrksich, Maleimide-Functionalized Self-Assembled

Monolayers for the Preparation of Peptide and Carbohydrate Biochips†, Langmuir, 19 (2002)

SC
1522-1531.

NU
[14] P. Jonkheijm, D. Weinrich, H. Schröder, C.M. Niemeyer, H. Waldmann, Chemical strategies

for generating protein biochips, Angew. Chem. Int. Ed., 47 (2008) 9618-9647.
MA
[15] A. Misra, P. Dwivedi, Immobilization of oligonucleotides on glass surface using an efficient

heterobifunctional reagent through maleimide–thiol combination chemistry, Anal. Biochem., 369


D

(2007) 248-255.
TE

[16] E.A. Smith, M.J. Wanat, Y. Cheng, S.V.P. Barreira, A.G. Frutos, R.M. Corn, Formation,
P

Spectroscopic Characterization, and Application of Sulfhydryl-Terminated Alkanethiol


CE

Monolayers for the Chemical Attachment of DNA onto Gold Surfaces, Langmuir, 17 (2001)
AC

2502-2507.

[17] H. Biederman, Introducion, in: H. Biederman (Ed.) Plasma Polymer Films, Imperial College

Press, London, 2004, pp. 15-16.

[18] N. Bondt, J.R. Deimann, A.P.v. Trostwijk, cited in J. Fourcroy Ann. Chem., 21 (1796) 58.

[19] J. Goodman, The formation of thin polymer films in the gas discharge, Journal of Polymer

Science, 44 (1960) 551-552.

[20] J. Friedrich, The Plasma Chemistry of Polymer Surfaces, Wiley, Weinhein, 2012.

[21] J. Friedrich, Mechanisms of Plasma Polymerization – Reviewed from a Chemical Point of

View, Plasma Process. Polym., 8 (2011) 783-802.


87
ACCEPTED MANUSCRIPT

[22] N. Inagaki, Plasma Surface Modification and Plasma Polymerization, Technomic

Publishing, Lancaster, 1996.

[23] L. Jorge, S. Coulombe, P.-L. Girard-Lauriault, Nanofluids Containing MWCNTs Coated

PT
with Nitrogen-Rich Plasma Polymer Films for CO2 Absorption in Aqueous Medium, Plasma

RI
Process. Polym., 12 (2015) 1311-1321.

[24] D. Shi, J. Lian, P. He, L. Wang, W.J. van Ooij, M. Schulz, Y. Liu, D.B. Mast, Plasma

SC
deposition of ultrathin polymer films on carbon nanotubes, Appl. Phys. Lett., 81 (2002) 5216-

NU
5218.

[25] A. Felten, C. Bittencourt, J.-J. Pireaux, G. Van Lier, J.-C. Charlier, Radio-frequency plasma
MA
functionalization of carbon nanotubes surface O2, NH3, and CF4 treatments, J. Appl. Phys., 98

(2005) 074308.
D

[26] Q. Chen, L. Dai, M. Gao, S. Huang, A. Mau, Plasma activation of carbon nanotubes for
TE

chemical modification, J. Phys. Chem. B, 105 (2001) 618-622.


P

[27] B. Akhavan, K. Jarvis, P. Majewski, Plasma polymerization of sulfur-rich and water-stable


CE

coatings on silica particles, Surf. Coat. Technol., 264 (2015) 72-79.


AC

[28] B. Akhavan, K. Jarvis, P. Majewski, Plasma Polymer-Functionalized Silica Particles for

Heavy Metals Removal, ACS Applied Materials & Interfaces, 7 (2015) 4265-4274.

[29] T.D. Michl, B.R. Coad, A. Hüsler, K. Vasilev, H.J. Griesser, Laboratory scale systems for

the plasma treatment and coating of particles, Plasma Process. Polym., 12 (2015) 305-313.

[30] K.L. Jarvis, P. Majewski, Influence of Particle Mass and Flow Rate on Plasma Polymerized

Allylamine Coated Quartz Particles for Humic Acid Removal, Plasma Process. Polym., 12 (2015)

42-50.

88
ACCEPTED MANUSCRIPT

[31] A. Choukourov, I. Melnichuk, A. Shelemin, P. Solař, J. Hanus, D. Slavinska, H. Biederman,

Plasma Polymerization on Mesoporous Surfaces: N-Hexane on Titanium Nano-Particles, J. Phys.

Chem. C, DOI 10.1021/acs.jpcc.5b08604(2015).

PT
[32] V. Barranco, J. Carpentier, G. Grundmeier, Correlation of morphology and barrier properties

RI
of thin microwave plasma polymer films on metal substrate, Electrochim. Acta, 49 (2004) 1999-

2013.

SC
[33] G. Grundmeier, P. Thiemann, J. Carpentier, V. Barranco, Tailored thin plasma polymers for

NU
the corrosion protection of metals, Surf. Coat. Technol., 174–175 (2003) 996-1001.

[34] M. Deilmann, S. Theiß, P. Awakowicz, Pulsed microwave plasma polymerization of silicon


MA
oxide films: Application of efficient permeation barriers on polyethylene terephthalate, Surf.

Coat. Technol., 202 (2008) 1911-1917.


D

[35] J. Schneider, D. Kiesler, M. Leins, A. Schulz, M. Walker, U. Schumacher, U. Stroth,


TE

Development of Plasma Polymerised SiOx Barriers on Polymer Films for Food Packaging
P

Applications, Plasma Process. Polym., 4 (2007) S155-S159.


CE

[36] M.R. Alexander, T.M. Duc, A study of the interaction of acrylic acid/1,7-octadiene plasma
AC

deposits with water and other solvents, Polymer, 40 (1999) 5479-5488.

[37] S. Fraser, R.D. Short, D. Barton, J.W. Bradley, A Multi-Technique Investigation of the

Pulsed Plasma and Plasma Polymers of Acrylic Acid:  Millisecond Pulse Regime, J. Phys.

Chem. B, 106 (2002) 5596-5603.

[38] D.B. Haddow, R.M. France, R.D. Short, J.W. Bradley, D. Barton, A Mass Spectrometric and

Ion Energy Study of the Continuous Wave Plasma Polymerization of Acrylic Acid, Langmuir, 16

(2000) 5654-5660.

[39] D. Hegemann, E. Körner, K. Albrecht, U. Schütz, S. Guimond, Growth Mechanism of

Oxygen-Containing Functional Plasma Polymers, Plasma Process. Polym., 7 (2010) 889-898.


89
ACCEPTED MANUSCRIPT

[40] D. Hegemann, U. Schütz, E. Körner, Macroscopic Approach to Plasma Polymerization

Using the Concept of Energy Density, Plasma Process. Polym., 8 (2011) 689-694.

[41] S.A. Voronin, M. Zelzer, C. Fotea, M.R. Alexander, J.W. Bradley, Pulsed and Continuous

PT
Wave Acrylic Acid Radio Frequency Plasma Deposits:  Plasma and Surface Chemistry, J. Phys.

RI
Chem. B, 111 (2007) 3419-3429.

[42] C.L. Rinsch, X. Chen, V. Panchalingam, R.C. Eberhart, J.-H. Wang, R.B. Timmons, Pulsed

SC
Radio Frequency Plasma Polymerization of Allyl Alcohol:  Controlled Deposition of Surface

NU
Hydroxyl Groups, Langmuir, 12 (1996) 2995-3002.

[43] L. Watkins, A. Bismarck, A.F. Lee, D. Wilson, K. Wilson, An XPS study of pulsed plasma
MA
polymerised allyl alcohol film growth on polyurethane, Appl. Surf. Sci., 252 (2006) 8203-8211.

[44] A. Choukourov, H. Biederman, I. Kholodkov, D. Slavinska, M. Trchova, A. Hollander,


D

Properties of amine-containing coatings prepared by plasma polymerization, J. Appl. Polym. Sci.,


TE

92 (2004) 979-990.
P

[45] A. Choukourov, H. Biederman, D. Slavinska, L. Hanley, A. Grinevich, H. Boldyryeva, A.


CE

Mackova, Mechanistic Studies of Plasma Polymerization of Allylamine, J. Phys. Chem. B, 109


AC

(2005) 23086-23095.

[46] D. Debarnot, T. Mérian, F. Poncin-Epaillard, Film Chemistry Control and Growth Kinetics

of Pulsed Plasma-Polymerized Aniline, Plasma Chem. Plasma Process., 31 (2010) 217-231.

[47] L. Denis, D. Cossement, T. Godfroid, F. Renaux, C. Bittencourt, R. Snyders, M. Hecq,

Synthesis of Allylamine Plasma Polymer Films: Correlation between Plasma Diagnostic and Film

Characteristics, Plasma Process. Polym., 6 (2009) 199-208.

[48] L. Denis, P. Marsal, Y. Olivier, T. Godfroid, R. Lazzaroni, M. Hecq, J. Cornil, R. Snyders,

Deposition of Functional Organic Thin Films by Pulsed Plasma Polymerization: A Joint

Theoretical and Experimental Study, Plasma Process. Polym., 7 (2010) 172-181.


90
ACCEPTED MANUSCRIPT

[49] K. Vasilev, L. Britcher, A. Casanal, H.J. Griesser, Solvent-Induced Porosity in Ultrathin

Amine Plasma Polymer Coatings, J. Phys. Chem. B, 112 (2008) 10915-10921.

[50] C. Daunton, L.E. Smith, J.D. Whittle, R.D. Short, D.A. Steele, A. Michelmore, Plasma

PT
Parameter Aspects in the Fabrication of Stable Amine Functionalized Plasma Polymer Films,

RI
Plasma Process. Polym., 12 (2015) 817-826.

[51] L. Denis, D. Thiry, D. Cossement, P. Gerbaux, F. Brusciotti, I. Van De Keere, V. Goossens,

SC
H. Terryn, M. Hecq, R. Snyders, Towards the understanding of plasma polymer film behaviour in

NU
ethanol: A multi-technique investigation, Prog. Org. Coat., 70 (2011) 134-141.

[52] S. Ligot, F. Renaux, L. Denis, D. Cossement, N. Nuns, P. Dubois, R. Snyders, Experimental


MA
Study of the Plasma Polymerization of Ethyl Lactate, Plasma Process. Polym., 10 (2013) 999-

1009.
D

[53] S. Ligot, M. Guillaume, P. Raynaud, D. Thiry, V. Lemaur, T. Silva, N. Britun, J. Cornil, P.


TE

Dubois, R. Snyders, Experimental and Theoretical Study of the Plasma Chemistry of Ethyl
P

Lactate Plasma Polymerization Discharges, Plasma Process. Polym., 12 (2015) 405-415.


CE

[54] S. Ligot, M. Guillaume, P. Gerbaux, D. Thiry, F. Renaux, J. Cornil, P. Dubois, R. Snyders,


AC

Combining Mass Spectrometry Diagnostic and Density Functional Theory Calculations for a

Better Understanding of the Plasma Polymerization of Ethyl Lactate, J. Phys. Chem. B, 118

(2014) 4201-4211.

[55] L. Denis, F. Renaux, D. Cossement, C. Bittencourt, N. Tuccitto, A. Licciardello, M. Hecq,

R. Snyders, Physico-Chemical Characterization of Methyl Isobutyrate-based Plasma Polymer

Films, Plasma Process. Polym., 8 (2011) 127-137.

[56] T.R. Gengenbach, H.J. Griesser, Deposition conditions influence the postdeposition

oxidation of methyl methacrylate plasma polymer films, J. Polym. Sci., Part A: Polym. Chem., 36

(1998) 985-1000.
91
ACCEPTED MANUSCRIPT

[57] A. Michelmore, P. Gross-Kosche, S.A. Al-Bataineh, J.D. Whittle, R.D. Short, On the Effect

of Monomer Chemistry on Growth Mechanisms of Nonfouling PEG-like Plasma Polymers,

Langmuir, 29 (2013) 2595-2601.

PT
[58] A. Michelmore, D.A. Steele, D.E. Robinson, J.D. Whittle, R.D. Short, The link between

RI
mechanisms of deposition and the physico-chemical properties of plasma polymer films, Soft

Matter, 9 (2013) 6167-6175.

SC
[59] G.P. Wells, I.C. Estrada-Raygoza, P.L.S. Thamban, C.T. Nelson, C.-W. Chung, L.J.

NU
Overzet, M.J. Goeckner, Understanding the Synthesis of Ethylene Glycol Pulsed Plasma

Discharges, Plasma Process. Polym., 10 (2013) 119-135.


MA
[60] R. D'Agostino, F. Cramarossa, F. Fracassi, E. Desimoni, L. Sabbatini, P.G. Zambonin, G.

Caporiccio, Polymer film formation in C2F6-H2 discharges, Thin Solid Films, 143 (1986) 163-
D

175.
TE

[61] P. Favia, Plasma deposition of fluoropolymer films in different glow discharges regimes, in:
P

H. Bierderman (Ed.) Plasma Polymer Films, Imperial College Press, London, 2004, pp. 46-47.
CE

[62] F. Henry, F. Renaux, S. Coppée, R. Lazzaroni, N. Vandencasteele, F. Reniers, R. Snyders,


AC

Synthesis of superhydrophobic PTFE-like thin films by self-nanostructuration in a hybrid plasma

process, Surf. Sci., 606 (2012) 1825-1829.

[63] R.T. Chen, B.W. Muir, G.K. Such, A. Postma, R.A. Evans, S.M. Pereira, K.M. McLean, F.

Caruso, Surface “Click” Chemistry on Brominated Plasma Polymer Thin Films, Langmuir, 26

(2009) 3388-3393.

[64] W.C.E. Schofield, J. McGettrick, T.J. Bradley, J.P.S. Badyal, S. Przyborski, Rewritable

DNA Microarrays, J. Am. Chem. Soc., 128 (2006) 2280-2285.

[65] L.G. Harris, W.C.E. Schofield, K.J. Doores, B.G. Davis, J.P.S. Badyal, Rewritable

Glycochips, J. Am. Chem. Soc., 131 (2009) 7755-7761.


92
ACCEPTED MANUSCRIPT

[66] D. Thiry, N. Britun, S. Konstantinidis, J.-P. Dauchot, M. Guillaume, J. Cornil, R. Snyders,

Experimental and Theoretical Study of the Effect of the Inductive-to-Capacitive Transition in

Propanethiol Plasma Polymer Chemistry, J. Phys. Chem. C, 117 (2013) 9843-9851.

PT
[67] D. Thiry, R. Francq, D. Cossement, D. Guerin, D. Vuillaume, R. Snyders, Establishment of a

RI
Derivatization Method To Quantify Thiol Function in Sulfur-Containing Plasma Polymer Films,

Langmuir, 29 (2013) 13183-13189.

SC
[68] D. Thiry, R. Francq, D. Cossement, M. Guillaume, J. Cornil, R. Snyders, A Detailed

NU
Description of the Chemistry of Thiol Supporting Plasma Polymer Films, Plasma Process.

Polym., 11 (2014) 606-615.


MA
[69] L.M.H. Groenewoud, G.H.M. Engbers, J. Feijen, Plasma Polymerization of Thiophene

Derivatives, Langmuir, 19 (2003) 1368-1374.


D

[70] L.M.H. Groenewoud, G.H.M. Engbers, J.G.A. Terlingen, H. Wormeester, J. Feijen, Pulsed
TE

Plasma Polymerization of Thiophene, Langmuir, 16 (2000) 6278-6286.


P

[71] S. Liu, M.M.L.M. Vareiro, S. Fraser, A.T.A. Jenkins, Control of Attachment of Bovine
CE

Serum Albumin to Pulse Plasma-Polymerized Maleic Anhydride by Variation of Pulse


AC

Conditions, Langmuir, 21 (2005) 8572-8575.

[72] K.S. Siow, L. Britcher, S. Kumar, H.J. Griesser, Plasma Methods for the Generation of

Chemically Reactive Surfaces for Biomolecule Immobilization and Cell Colonization - A

Review, Plasma Process. Polym., 3 (2006) 392-418.

[73] M.N. Macgregor-Ramiasa, A.A. Cavallaro, K. Vasilev, Properties and reactivity of

polyoxazoline plasma polymer films, Journal of Materials Chemistry B, 3 (2015) 6327-6337.

[74] P. Qi, W. Yan, Y. Yang, Y. Li, Y. Fan, J. Chen, Z. Yang, Q. Tu, N. Huang, Immobilization

of DNA aptamers via plasma polymerized allylamine film to construct an endothelial progenitor

cell-capture surface, Colloids and Surfaces B: Biointerfaces, 126 (2015) 70-79.


93
ACCEPTED MANUSCRIPT

[75] B.R. Coad, M. Jasieniak, S.S. Griesser, H.J. Griesser, Controlled covalent surface

immobilisation of proteins and peptides using plasma methods, Surf. Coat. Technol., 233 (2013)

169-177.

PT
[76] A. Manakhov, P. Skládal, D. Nečas, J. Čechal, J. Polčák, M. Eliáš, L. Zajíčková,

RI
Cyclopropylamine plasma polymers deposited onto quartz crystal microbalance for biosensing

application, physica status solidi (a), 211 (2014) 2801-2808.

SC
[77] K. Bazaka, M. Jacob, W. Chrzanowski, K. Ostrikov, Anti-bacterial surfaces: natural agents,

NU
mechanisms of action, and plasma surface modification, RSC Adv., 5 (2015) 48739-48759.

[78] S. Bhatt, J. Pulpytel, F. Arefi-Khonsari, Low and atmospheric plasma polymerisation of


MA
nanocoatings for bio-applications, Surface Innovations, 3 (2015) 63-83.

[79] P. Qi, Y. Yang, K.-Q. Xiong, Q. Tu, Z. Yang, J. Wang, J. Chen, N. Huang, J. Wang,
D

Multifunctional Plasma Polymerized Film: towards better anti-corrosion property, enhanced


TE

cellular growth ability, attenuated inflammatory and histological responses, ACS Biomater. Sci.
P

Eng., 1 (2015) 513-524.


CE

[80] J.W. Haycock, 3D cell culture: a review of current approaches and techniques, in: J.W.
AC

Haycock (Ed.) 3D Cell Culture, Springer, 2011, pp. 1-15.

[81] J.F. Friedrich, R. Mix, G. Kühn, Functional groups bearing plasma homo and copolymer

layers as adhesion promoters in metal–polymer composites, Surf. Coat. Technol., 174–175

(2003) 811-815.

[82] R. Förch, Z. Zhang, W. Knoll, Soft Plasma Treated Surfaces: Tailoring of Structure and

Properties for Biomaterial Applications, Plasma Process. Polym., 2 (2005) 351-372.

[83] S. Taheri, A. Cavallaro, S. Christo, P.J. Majewski, M. Barton, J.D. Hayball, K. Vasilev,

Antibacterial plasma polymer films conjugated with phospholipid encapsulated silver

nanoparticles, ACS Biomater. Sci. Eng., 1 (2015) 1278-1286.


94
ACCEPTED MANUSCRIPT

[84] S. Bhatt, J. Pulpytel, M. Mirshahi, F. Arefi-Khonsari, Plasma co-polymerized nano coatings

– As a biodegradable solid carrier for tunable drug delivery applications, Polymer, 54 (2013)

4820-4829.

PT
[85] K. Vasilev, N. Poulter, P. Martinek, H.J. Griesser, Controlled Release of Levofloxacin

RI
Sandwiched between Two Plasma Polymerized Layers on a Solid Carrier, ACS Applied

Materials & Interfaces, 3 (2011) 4831-4836.

SC
[86] K. Vasilev, Z. Poh, K. Kant, J. Chan, A. Michelmore, D. Losic, Tailoring the surface

NU
functionalities of titania nanotube arrays, Biomaterials, 31 (2010) 532-540.

[87] S. Simovic, D. Losic, K. Vasilev, Controlled drug release from porous materials by plasma
MA
polymer deposition, Chem. Commun., 46 (2010) 1317-1319.

[88] O. Kylián, A. Choukourov, H. Biederman, Nanostructured plasma polymers, Thin Solid


D

Films, 548 (2013) 1-17.


TE

[89] P.L. Girard‐Lauriault, W.E. Unger, P.M. Dietrich, A. Holländer, Innovative and Established
P

Strategies for the Surface Analysis of Nitrogen and Oxygen‐Rich Plasma Polymer Films by XPS:
CE

An Introductory Guide, Plasma Process. Polym., 12 (2015) 953-967.


AC

[90] B. Nisol, F. Reniers, Challenges in the characterization of plasma polymers using XPS, J.

Electron. Spectrosc. Relat. Phenom., 200 (2015) 311-331.

[91] A. Bogaerts, E. Neyts, R. Gijbels, J. van der Mullen, Gas discharge plasmas and their

applications, Spectrochimica Acta Part B: Atomic Spectroscopy, 57 (2002) 609-658.

[92] M.A. Lieberman, A.J. Lichtenberg, Principle of Plasma Discharge and Materials Processing,

Wiley, New York, 2005.

[93] A. Fridman, Plasma Chemistry, Cambridge University Press, New York, 2008.

[94] D. Merche, N. Vandencasteele, F. Reniers, Atmospheric plasmas for thin film deposition: A

critical review, Thin Solid Films, 520 (2012) 4219-4236.


95
ACCEPTED MANUSCRIPT

[95] S. Guimond, U. Schütz, B. Hanselmann, E. Körner, D. Hegemann, Influence of gas phase

and surface reactions on plasma polymerization, Surf. Coat. Technol., 205, Supplement 2 (2011)

S447-S450.

PT
[96] H. Yasuda, Luminous Chemical Vapor Deposition And Interface, Marcel Dekker, New

RI
York, 2005.

[97] H. Yasuda, Plasma Polymerization, Elsevier Science, 1985.

SC
[98] A. Milella, F. Palumbo, P. Favia, G. Cicala, R. d'Agostino, Continuous and Modulated

NU
Deposition of Fluorocarbon Films from c-C4F8 Plasmas, Plasma Process. Polym., 1 (2004) 164-

170.
MA
[99] S. Ligot, D. Thiry, P.A. Cormier, P. Raynaud, P. Dubois, R. Snyders, In situ IR

Spectroscopy as a Tool to Better Understand the Growth Mechanisms of Plasma Polymers Thin
D

Films, Plasma Process. Polym., 12 (2015) 1200-1207.


TE

[100] A. Michelmore, J.D. Whittle, R.D. Short, The importance of ions in low pressure PECVD
P

plasmas, Frontiers in Physics, 3 (2015) 3.


CE

[101] S. Candan, A. J. Beck, L. O'Toole, R. D. Short, A. Goodyear, N. St J. Braithwaite, The role


AC

of ions in the continuous-wave plasma polymerisation of acrylic acid, PCCP, 1 (1999) 3117-

3121.

[102] A. Michelmore, P.M. Bryant, D.A. Steele, K. Vasilev, J.W. Bradley, R.D. Short, Role of

Positive Ions in Determining the Deposition Rate and Film Chemistry of Continuous Wave

Hexamethyl Disiloxane Plasmas, Langmuir, 27 (2011) 11943-11950.

[103] A. Michelmore, C. Charles, R.W. Boswell, R.D. Short, J.D. Whittle, Defining Plasma

Polymerization: New Insight Into What We Should Be Measuring, ACS Applied Materials &

Interfaces, 5 (2013) 5387-5391.

96
ACCEPTED MANUSCRIPT

[104] A.J. Beck, S. Candan, R.D. Short, A. Goodyear, N.S.J. Braithwaite, The Role of Ions in the

Plasma Polymerization of Allylamine, J. Phys. Chem. B, 105 (2001) 5730-5736.

[105] P.N. Brookes, S. Fraser, R.D. Short, L. Hanley, E. Fuoco, A. Roberts, S. Hutton, The effect

PT
of ion energy on the chemistry of air-aged polymer films grown from the hyperthermal

polyatomic ion Si2OMe5+, J. Electron. Spectrosc. Relat. Phenom., 121 (2001) 281-297.

RI
[106] A. Choukourov, J. ousal, D. Slav nská, H. Biederman, E.R. Fuoco, S. Tepavcevic, J.

SC
Saucedo, L. Hanley, Growth of primary and secondary amine films from polyatomic ion

NU
deposition, Vacuum, 75 (2004) 195-205.

[107] A. Von Keudell, T. Schwarz-Selinger, M. Meier, W. Jacob, Direct identification of the


MA
synergism between methyl radicals and atomic hydrogen during growth of amorphous

hydrogenated carbon films, Appl. Phys. Lett., 76 (2000) 676-678.


D

[108] P. Traskelin, O. Saresoja, K. Nordlund, Molecular dynamics simulations of C2, C2H, C2H2,
TE

C2H3, C2H4, C2H5, and C2H6 bombardment of diamond (111) surfaces, J. Nucl. Mater., 375
P

(2008) 270-274.
CE

[109] A. Von Keudell, Surface Processes during thin-film growth, Plasma Sources Sci. Technol.,
AC

9 (2000) 455-467.

[110] D. Thiry, A. De Vreese, F. Renaux, J.L. Colaux, S. Lucas, Y. Guinet, L. Paccou, E.

Bousser, R. Snyders, Toward a Better Understanding of the Influence of the Hydrocarbon

Precursor on the Mechanical Properties of a-C:H Coatings Synthesized by a Hybrid

PECVD/PVD Method, Plasma Process. Polym., DOI 10.1002/ppap.201500050(2015).

[111] J. Robertson, Diamond-like amorphous carbon, Materials Science and Engineering: R:

Reports, 37 (2002) 129-281.

97
ACCEPTED MANUSCRIPT

[112] J.D. Whittle, D.A. Steele, R.D. Short, Reconciling the Physical and Chemical

Environments of Plasma: A Commentary on “Mechanisms of Plasma Polymerisation–Reviewed

from a Chemical Point of View”, Plasma Process. Polym., 9 (2012) 840-843.

PT
[113] R. d'Agostino, F. Palumbo, Comment on “Ion‐Assisted Processes of Polymerization in

Low‐Pressure Plasmas”, Plasma Process. Polym., 9 (2012) 844-849.

RI
[114] T.R. Gengenbach, Z.R. Vasic, R.C. Chatelier, H.J. Griesser, A multi-technique study of the

SC
spontaneous oxidation of N-hexane plasma polymers, J. Polym. Sci., Part A: Polym. Chem., 32

NU
(1994) 1399-1414.

[115] R.C. Chatelier, X. Xie, T.R. Gengenbach, H.J. Griesser, Quantitative Analysis of Polymer
MA
Surface Restructuring, Langmuir, 11 (1995) 2576-2584.

[116] J.D. Whittle, R.D. Short, C.W.I. Douglas, J. Davies, Differences in the Aging of Allyl
D
TE

Alcohol, Acrylic Acid, Allylamine, and Octa-1,7-diene Plasma Polymers As Studied by X-ray

Photoelectron Spectroscopy, Chem. Mater., 12 (2000) 2664-2671.


P

[117] S. Ershov, F. Khelifa, P. Dubois, R. Snyders, Derivatization of Free Radicals in an


CE

Isopropanol Plasma Polymer Film: The First Step toward Polymer Grafting, ACS Applied
AC

Materials & Interfaces, 5 (2013) 4216-4223.

[118] P.-L. Girard-Lauriault, P.M. Dietrich, T. Gross, T. Wirth, W.E.S. Unger, Chemical

Characterization of the Long-Term Ageing of Nitrogen-Rich Plasma Polymer Films under

Various Ambient Conditions, Plasma Process. Polym., 10 (2013) 388-395.

[119] T.R. Gengenbach, H.J. Griesser, Aging of 1,3-diaminopropane plasma-deposited polymer

films: Mechanisms and reaction pathways, J. Polym. Sci., Part A: Polym. Chem., 37 (1999)

2191-2206.

98
ACCEPTED MANUSCRIPT

[120] S. Ershov, F. Khelifa, M.-E. Druart, Y. Habibi, M.-G. Olivier, R. Snyders, P. Dubois, Free

radical-induced grafting from plasma polymers for the synthesis of thin barrier coatings, RSC

Adv., 5 (2015) 14256-14265.

PT
[121] F. Khelifa, S. Ershov, M.-E. Druart, Y. Habibi, D. Chicot, M.-G. Olivier, R. Snyders, P.

RI
Dubois, A multilayer coating with optimized properties for corrosion protection of Al, J. Mater.

Chem. A, 3 (2015) 15977-15985.

SC
[122] D. Teare, W. Schofield, R. Garrod, J. Badyal, Rapid polymer brush growth by TEMPO-

NU
mediated controlled free-radical polymerization from swollen plasma deposited poly (maleic

anhydride) initiator surfaces, Langmuir, 21 (2005) 10818-10824.


MA
[123] D. Teare, W. Schofield, V. Roucoules, J. Badyal, Substrate-independent growth of

micropatterned polymer brushes, Langmuir, 19 (2003) 2398-2403.


D

[124] F. Bénard, P. Dubois, M. Olivier, R. Snyders, L. Denis, F. Khelifa, D. Thiry, F. Renaux,


TE

Grafted polymer coatings, University of Mons and Materia Nova, WO2011092212A1 (2011).
P

[125] Y. Yin, K. Fisher, N.J. Nosworthy, D. Bax, R.J. Clarke, D.R. McKenzie, M.M. Bilek,
CE

Comparison on protein adsorption properties of diamond-like carbon and nitrogen-containing


AC

plasma polymer surfaces, Thin Solid Films, 520 (2012) 3021-3025.

[126] Y. Yin, M.M. Bilek, D.R. McKenzie, Direct Evidence of Covalent Immobilisation of

Microperoxidase‐11 on Plasma Polymer Surfaces, Plasma Process. Polym., 7 (2010) 708-714.

[127] L.-Q. Chu, W. Knoll, R. Förch, Stabilization of Plasma-Polymerized Allylamine Films by

Ethanol Extraction, Langmuir, 22 (2006) 5548-5551.

[128] S. Candan, A.J. Beck, L. Otoole, R.D. Short, Effects of "processing parameters" in plasma

deposition: Acrylic acid revisited, J. Vac. Sci. Technol., A, 16 (1998) 1702-1709.

99
ACCEPTED MANUSCRIPT

[129] D.B. Haddow, A. Goruppa, J. Whittle, R.D. Short, O. Kahle, C. Uhlig, M. Bauer,

Application of Variable-Temperature Ellipsometry to Plasma Polymers:  The Effect of Addition

of 1,7-Octadiene to Plasma Deposits of Acrylic Acid, Chem. Mater., 12 (2000) 866-868.

PT
[130] S. Swaraj, U. Oran, A. Lippitz, J.F. Friedrich, W.E.S. Unger, Aging of Plasma-Deposited

RI
Films Prepared from Organic Monomers, Plasma Process. Polym., 4 (2007) S784-S789.

[131] D. Thiry, F.J. Aparicio, N. Britun, R. Snyders, Concomitant effects of the substrate

SC
temperature and the plasma chemistry on the chemical properties of propanethiol plasma polymer

NU
prepared by ICP discharges, Surf. Coat. Technol., 241 (2014) 2-7.

[132] D. Thiry, N. Britun, S. Konstantinidis, J.-P. Dauchot, L. Denis, R. Snyders, Altering the
MA
sulfur content in the propanethiol plasma polymers using the capacitive-to-inductive mode

transition in inductively coupled plasma discharge, Appl. Phys. Lett., 100 (2012) 071604.
D

[133] B. Finke, K. Schröder, A. Ohl, Structure Retention and Water Stability of Microwave
TE

Plasma Polymerized Films From Allylamine and Acrylic Acid, Plasma Process. Polym., 6 (2009)
P

S70-S74.
CE

[134] D.E. Robinson, D.J. Buttle, J.D. Whittle, K.L. Parry, R.D. Short, D.A. Steele, The
AC

Substrate and Composition Dependence of Plasma Polymer Stability, Plasma Process. Polym., 7

(2010) 102-106.

[135] S. Swaraj, U. Oran, A. Lippitz, J.F. Friedrich, W.E.S. Unger, Surface Analysis of Plasma-

Deposited Polymer Films, 6, Plasma Process. Polym., 2 (2005) 572-580.

[136] Q. Chen, R. Förch, W. Knoll, Characterization of Pulsed Plasma Polymerization

Allylamine as an Adhesion Layer for DNA Adsorption/Hybridization, Chem. Mater., 16 (2004)

614-620.

[137] Z. Zhang, Q. Chen, W. Knoll, R. Foerch, R. Holcomb, D. Roitman, Plasma Polymer Film

Structure and DNA Probe Immobilization, Macromolecules, 36 (2003) 7689-7694.


100
ACCEPTED MANUSCRIPT

[138] D. Szmigiel, C. Hibert, A. Bertsch, E. Pamuła, . Domański, P. Grabiec, P. Prokaryn, A.

Ścisłowska-Czarnecka, B. Płytycz, Fluorine-Based Plasma Treatment of Biocompatible Silicone

Elastomer: The Effect of Temperature on Etch Rate and Surface Properties, Plasma Process.

PT
Polym., 5 (2008) 246-255.

RI
[139] T.B. Casserly, K.K. Gleason, Effect of Substrate Temperature on the Plasma

Polymerization of Poly(methyl methacrylate), Chem. Vap. Deposition, 12 (2006) 59-66.

SC
[140] D. Cossement, F. Renaux, D. Thiry, S. Ligot, R. Francq, R. Snyders, Chemical and

NU
microstructural characterizations of plasma polymer films by time-of-flight secondary ion mass

spectrometry and principal component analysis, Appl. Surf. Sci., 355 (2015) 842-848.
MA
[141] J.D. Whittle, R.D. Short, D.A. Steele, J.W. Bradley, P.M. Bryant, F. Jan, H. Biederman,

A.A. Serov, A. Choukurov, A.L. Hook, Variability in Plasma Polymerization Processes–An


D

International Round‐Robin Study, Plasma Process. Polym., 10 (2013) 767-778.


TE

[142] M. Dhayal, J.W. Bradley, Using heated probes in plasma polymerising discharges, Surf.
P

Coat. Technol., 184 (2004) 116-122.


CE

[143] M.A. Gilliam, Q. Yu, H. Yasuda, Plasma Polymerization Behavior of Fluorocarbon


AC

Monomers in Low-Pressure AF and RF Discharges, Plasma Process. Polym., 4 (2007) 165-172.

[144] D. Hegemann, Macroscopic investigation of reaction rates yielding plasma polymer

deposition, J. Phys. D: Appl. Phys., 46 (2013) 205204.

[145] D. Hegemann, Plasma polymerization and its applications in textiles, INDIAN JOURNAL

OF FIBRE AND TEXTILE RESEARCH, 31 (2006) 99.

[146] D. Hegemann, M.M. Hossain, E. Körner, D.J. Balazs, Macroscopic Description of Plasma

Polymerization, Plasma Process. Polym., 4 (2007) 229-238.

[147] D. Hegemann, H. Brunner, C. Oehr, Deposition rate and three-dimensional uniformity of

RF plasma deposited SiO x films, Surf. Coat. Technol., 142 (2001) 849-855.
101
ACCEPTED MANUSCRIPT

[148] D. Hegemann, E. Koerner, S. Guimond, Plasma polymerization of acrylic acid revisited,

Plasma Process. Polym., 6 (2009) 246-254.

[149] D. Hegemann, Macroscopic control of plasma polymerization processes, Pure Appl.

PT
Chem., 80 (2008) 1893-1900.

RI
[150] D. Hegemann, M.M. Hossain, Influence of Non‐Polymerizable Gases Added During

Plasma Polymerization, Plasma Process. Polym., 2 (2005) 554-562.

SC
[151] D. Hegemann, E. örner, S. Guimond, Reply to: “Testing the Hypothesis: Comments on

NU
Plasma Polymerization of Acrylic Acid Revisited”, Plasma Process. Polym., 7 (2010) 371-375.

[152] R.D. Short, D.A. Steele, Testing the hypothesis: comments on plasma polymerisation of
MA
acrylic acid revisited, Plasma Process. Polym., 7 (2010) 366-370.

[153] H. Biederman, O. Kylián, Some Remarks to Macroscopic Kinetics of Plasma


D

Polymerization, Plasma Process. Polym., 8 (2011) 475-477.


TE

[154] A. von eudell, J. Benedikt, A Physicist's Perspective on “Views on Macroscopic inetics


P

of Plasma Polymerisation”, Plasma Process. Polym., 7 (2010) 376-379.


CE

[155] R.B. Timmons, A.J. Griggs, Plasma Polymer Films, in: H. Bierderman (Ed.) Plasma
AC

Polymer Films, Imperial College Press, London, 2004, pp. 217-245.

[156] R. Jafari, M. Tatoulian, W. Morscheidt, F. Arefi-Khonsari, Stable plasma polymerized

acrylic acid coating deposited on polyethylene (PE) films in a low frequency discharge (70kHz),

React. Funct. Polym., 66 (2006) 1757-1765.

[157] D. Hegemann, E. Körner, N. Blanchard, M. Drabik, S. Guimond, Densification of

functional plasma polymers by momentum transfer during film growth, Appl. Phys. Lett., 101

(2012) 211603.

102
ACCEPTED MANUSCRIPT

[158] D. Hegemann, B. Hanselmann, S. Guimond, G. Fortunato, M.-N. Giraud, A.G. Guex,

Considering the degradation effects of amino-functional plasma polymer coatings for biomedical

application, Surf. Coat. Technol., 255 (2014) 90-95.

PT
[159] A. Choukourov, H. Biederman, D. Slavinska, M. Trchova, A. Hollander, The influence of

RI
pulse parameters on film composition during pulsed plasma polymerization of

diaminocyclohexane, Surf. Coat. Technol., 174–175 (2003) 863-866.

SC
[160] T. Mérian, D. Debarnot, F. Poncin-Epaillard, Effect of Fluorine Substitution of Aniline

NU
Ring on Pulsed Plasma Polymer Growth and Structure, Plasma Process. Polym., 8 (2011) 763-

772.
MA
[161] G. Mishra, S.L. McArthur, Plasma Polymerization of Maleic Anhydride: Just What Are the

Right Deposition Conditions?, Langmuir, 26 (2010) 9645-9658.


D

[162] M. Dhayal, J.W. Bradley, Time-resolved electric probe measurements in the pulsed-plasma
TE

polymerisation of acrylic acid, Surf. Coat. Technol., 194 (2005) 167-174.


P

[163] S. Voronin, M. Alexander, J. Bradley, Time-resolved mass and energy spectral


CE

investigation of a pulsed polymerising plasma struck in acrylic acid, Surf. Coat. Technol., 201
AC

(2006) 768-775.

[164] M. Drabik, C. Celma, J. Kousal, H. Biederman, D. Hegemann, Properties of aC: H: O

plasma polymer films deposited from acetone vapors, Thin Solid Films, 573 (2014) 27-32.

[165] A. Bousquet, A. Granier, G. Cartry, A. Goullet, Kinetics of O and H atoms in pulsed

O2/HMDSO low pressure PECVD plasmas, Journal of optoelectronics and advanced materials,

10 (2008) 1999-2002.

[166] A. Bousquet, G. Cartry, A. Granier, Investigation of O-atom kinetics in O2, CO2, H2O and

O2/HMDSO low pressure radiofrequency pulsed plasmas by time-resolved optical emission

spectroscopy, Plasma Sources Sci. Technol., 16 (2007) 597.


103
ACCEPTED MANUSCRIPT

[167] F. Palumbo, P. Favia, A. Rinaldi, M. Vulpio, R. d'Agostino, PE-CVD of organic thin films

with controlled surface concentration of carboxylic groups, Plasmas Polym., 4 (1999) 133-145.

[168] A. Choudhury, J. Chutia, S. Barve, H. Kakati, A. Pal, N. Mithal, R. Kishore, M. Pandey, D.

PT
Patil, Studies of physical and chemical properties of styrene-based plasma polymer films

RI
deposited by radiofrequency Ar/styrene glow discharge, Prog. Org. Coat., 70 (2011) 75-82.

[169] L. O'Toole, R.D. Short, A.P. Ameen, F.R. Jones, Mass spectrometry of and deposition-rate

SC
measurements from radiofrequency-induced plasmas of methyl isobutyrate, methyl methacrylate

NU
and n-butyl methacrylate, J. Chem. Soc., Faraday Trans., 91 (1995) 1363-1370.

[170] A.J. Beck, S. Candan, R.M. France, F.R. Jones, R.D. Short, A mass spectral investigation
MA
of the RF plasmas of small organic compounds: An investigation of the plasma-phase reactions in

the plasma deposition from allyl amine, Plasmas Polym., 3 (1998) 97-114.
D

[171] J. Ahmad, K. Bazaka, J.D. Whittle, A. Michelmore, M.V. Jacob, Structural


TE

Characterization of γ-Terpinene Thin Films Using Mass Spectroscopy and X-Ray Photoelectron
P

Spectroscopy, Plasma Process. Polym., 12 (2015) 1085-1094.


CE

[172] H.D. Hazrati, J.D. Whittle, K. Vasilev, A mechanistic study of the plasma polymerization
AC

of ethanol, Plasma Process. Polym., 11 (2014) 149-157.

[173] H. Singh, J.W. Coburn, D.B. Graves, Mass spectrometric detection of reactive neutral

species: Beam-to-background ratio, Journal of Vacuum Science & Technology A, 17 (1999)

2447-2455.

[174] J. Benedikt, S. Agarwal, D. Eijkman, W. Vandamme, M. Creatore, M. Van de Sanden,

Threshold ionization mass spectrometry of reactive species in remote Ar∕ C2H2 expanding

thermal plasma, Journal of Vacuum Science & Technology A, 23 (2005) 1400-1412.

[175] J. Benedikt, A. Hecimovic, D. Ellerweg, A. Von Keudell, Quadrupole mass spectrometry

of reactive plasmas, J. Phys. D: Appl. Phys., 45 (2012) 403001.


104
ACCEPTED MANUSCRIPT

[176] N.S.J. Braithwaite, J. Booth, G. Cunge, A novel electrostatic probe method for ion flux

measurements, Plasma Sources Sci. Technol., 5 (1996) 677.

[177] P. Raynaud, T. Amilis, Y. Segui, Infrared absorption analysis of organosilicon/oxygen

PT
plasmas in a microwave multipolar plasma excited by distributed electron cyclotron resonance,

RI
Appl. Surf. Sci., 138 (1999) 285-291.

[178] P. Raynaud, B. Despax, Y. Segui, H. Caquineau, FTIR plasma phase analysis of

SC
hexamethyldisiloxane discharge in microwave multipolar plasma at different electrical powers,

NU
Plasma Process. Polym., 2 (2005) 45-52.

[179] A. Bogaerts, M. Eckert, M. Mao, E. Neyts, Computer modelling of the plasma chemistry
MA
and plasma-based growth mechanisms for nanostructured materials, J. Phys. D: Appl. Phys., 44

(2011) 174030.
D

[180] A.C. Van Duin, S. Dasgupta, F. Lorant, W.A. Goddard, ReaxFF: a reactive force field for
TE

hydrocarbons, J. Phys. Chem. A, 105 (2001) 9396-9409.


P

[181] R. Car, M. Parrinello, Unified approach for molecular dynamics and density-functional
CE

theory, Phys. Rev. Lett., 55 (1985) 2471.


AC

[182] A. Cheesman, J.N. Harvey, M.N. Ashfold, Studies of carbon incorporation on the diamond

{100} surface during chemical vapor deposition using density functional theory, J. Phys. Chem.

A, 112 (2008) 11436-11448.

[183] T.B. Casserly, K.K. Gleason, Chemical Vapor Deposition of Organosilicon Thin Films

from Methylmethoxysilanes, Plasma Process. Polym., 2 (2005) 679-687.

[184] K. Fukui, Formulation of the reaction coordinate, The Journal of Physical Chemistry, 74

(1970) 4161-4163.

105
ACCEPTED MANUSCRIPT

[185] E.C. Yang, X.J. Zhao, P. Tian, J.K. Hao, Density functional theory and MP2 calculations of

the transition states and reaction paths on coupling reaction of methane through plasma, Chin. J.

Chem . 22 (2004) 430-433.

PT
[186] L.A. Curtiss, K. Raghavachari, P.C. Redfern, J.A. Pople, Assessment of Gaussian-2 and

RI
density functional theories for the computation of enthalpies of formation, The Journal of

chemical physics, 106 (1997) 1063-1079.

SC
[187] O. Auciello, D.L. Flamm, Plasma Diagnostics: Surface Analysis and Interactions, Elsevier

NU
Science, San Diego, 2013.

[188] H. Biederman, P. Hlídek, J. Zemek, D. Slavínská, J. Jezek, P. Zakouril, J. Glosik,


MA
Deposition and properties of hydrophilic films prepared by plasma polymerization of Ar/n-

hexane/H2O, Vacuum, 46 (1995) 1413-1418.


D

[189] A. Brockhaus, G. Leu, V. Selenin, K. Tarnev, J. Engemann, Electron release in the


TE

afterglow of a pulsed inductively-coupled radiofrequency oxygen plasma, Plasma Sources Sci.


P

Technol., 15 (2006) 171.


CE

[190] J. Greene, Optical spectroscopy for diagnostics and process control during glow discharge
AC

etching and sputter deposition, J. Vac. Sci. Technol., 15 (1978) 1718-1729.

[191] J. Greene, F. Sequeda-Osorio, B. Natarajan, Glow discharge optical spectroscopy for

microvolume elemental analysis, J. Appl. Phys., 46 (1975) 2701-2709.

[192] A. Manakhov, L. Zajíčková, M. Eliáš, J. Čechal, J. Polčák, J. Hnilica, Š. Bittnerová, D.

Nečas, Optimization of Cyclopropylamine Plasma Polymerization toward Enhanced Layer

Stability in Contact with Water, Plasma Process. Polym., 11 (2014) 532-544.

[193] M. Buddhadasa, P.-L. Girard-Lauriault, Plasma co-polymerisation of ethylene, 1, 3-

butadiene and ammonia mixtures: Amine content and water stability, Thin Solid Films, 591

(2015) 76-85.
106
ACCEPTED MANUSCRIPT

[194] A. Choukourov, H. Biederman, I. Kholodkov, D. Slavinska, M. Trchova, A. Hollander,

Properties of amine‐containing coatings prepared by plasma polymerization, J. Appl. Polym. Sci.,

92 (2004) 979-990.

PT
[195] A. Choukourov, H. Biederman, D. Slavinska, M. Trchova, A. Hollander, The influence of

RI
pulse parameters on film composition during pulsed plasma polymerization of

diaminocyclohexane, Surf. Coat. Technol., 174 (2003) 863-866.

SC
[196] H. Yasuda, T. Yasuda, The competitive ablation and polymerization (CAP) principle and

NU
the plasma sensitivity of elements in plasma polymerization and treatment, J. Polym. Sci., Part A:

Polym. Chem., 38 (2000) 943-953.


MA
[197] R. d'Agostino, F. Cramarossa, S. De Benedictis, Diagnostics and decomposition

mechanism in radio-frequency discharges of fluorocarbons utilized for plasma etching or


D

polymerization, Plasma Chem. Plasma Process., 2 (1982) 213-231.


TE

[198] K. Aumaille, A. Granier, M. Schmidt, B. Grolleau, C. Vallee, G. Turban, Study of


P

oxygen/tetraethoxysilane plasmas in a helicon reactor using optical emission spectroscopy and


CE

mass spectrometry, Plasma Sources Sci. Technol., 9 (2000) 331.


AC

[199] A. Granier, M. Vervloet, K. Aumaille, C. Vallée, Optical emission spectra of TEOS and

HMDSO derived plasmas used for thin film deposition, Plasma Sources Sci. Technol., 12 (2003)

89.

[200] M. Goujon, T. Belmonte, G. Henrion, OES and FTIR diagnostics of HMDSO/O2 gas

mixtures for SiOx deposition assisted by RF plasma, Surf. Coat. Technol., 188 (2004) 756-761.

[201] N. Benissad, C. Boisse-Laporte, C. Vallée, A. Granier, A. Goullet, Silicon dioxide

deposition in a microwave plasma reactor, Surf. Coat. Technol., 116 (1999) 868-873.

107
ACCEPTED MANUSCRIPT

[202] J.-S. Yoon, M.-Y. Song, H. Kato, M. Hoshino, H. Tanaka, M.J. Brunger, S. Buckman, H.

Cho, Elastic cross sections for electron collisions with molecules relevant to plasma processing,

J. Phys. Chem. Ref. Data, 39 (2010) 033106.

PT
[203] J.-S. Yoon, M.-Y. Song, J.-M. Han, S.H. Hwang, W.-S. Chang, B. Lee, Y. Itikawa, Cross

RI
sections for electron collisions with hydrogen molecules, J. Phys. Chem. Ref. Data, 37 (2008)

913-931.

SC
[204] Y. Itikawa, A. Ichimura, K. Onda, K. Sakimoto, K. Takayanagi, Y. Hatano, M. Hayashi, H.

NU
Nishimura, S. Tsurubuchi, Cross sections for collisions of electrons and photons with oxygen

molecules, J. Phys. Chem. Ref. Data, 18 (1989) 23-42.


MA
[205] S.F. Durrant, R. Landers, G.G. Kleiman, S.G. Castro, M.A.B. de Moraes, Fluorine-

containing amorphous hydrogenated carbon films, Thin Solid Films, 281 (1996) 294-297.
D

[206] S.F. Durrant, M.A.B. de Moraes, Dynamic actinometric optical emission spectroscopy for
TE

the elucidation of plasma processes in the production of fluorinated amorphous hydrogenated


P

carbon films from glow discharges, Thin Solid Films, 277 (1996) 115-120.
CE

[207] S.F. Durrant, M.A.B. de Moraes, Conventional and dynamic actinometry of glow
AC

discharges fed mixtures of tetramethylsilane, sulfur hexafluoride, and helium, Journal of Vacuum

Science & Technology A, 16 (1998) 509-513.

[208] S.F. Durrant, M.A.B. De Moraes, PECVD of amorphous hydrogenated oxygenated

nitrogenated carbon films, J. Polym. Sci., Part B: Polym. Phys., 36 (1998) 1881-1888.

[209] A. Bouchoule, P. Ranson, Study of volume and surface processes in low pressure radio

frequency plasma reactors by pulsed excitation methods. I. Hydrogen–argon plasma, Journal of

Vacuum Science & Technology A, 9 (1991) 317-326.

[210] G. Turban, B. Grolleau, P. Launay, P. Briaud, A mass spectrometric diagnostic of C2F6 and

CHF3 plasmas during etching of SiO2 and Si, Revue de physique appliquée, 20 (1985) 609-620.
108
ACCEPTED MANUSCRIPT

[211] F. Moix, K. McKay, J.L. Walsh, J.W. Bradley, Atmospheric-Pressure Plasma

Polymerization of Acrylic Acid: Gas-Phase Ion Chemistry, Plasma Process. Polym., DOI

10.1002/ppap.201500031(2015).

PT
[212] E. Stoffels, W. Stoffels, K. Tachibana, Electron attachment mass spectrometry as a

RI
diagnostics for electronegative gases and plasmas, Rev. Sci. Instrum., 69 (1998) 116-122.

[213] E. de Hoffmann, V. Stroobant, Mass Spectrometry: Principles and Applications, Wiley,

SC
Chichester, 2013.

NU
[214] S. Candan, Radio Frequency-Induced Plasma Polymerization of Allyl Alcohol and 1-

Propanol, Turkish Journal of Chemistry, 26 (2002) 783-792.


MA
[215] L. O'Toole, A.J. Beck, A.P. Ameen, F.R. Jones, R.D. Short, Radiofrequency-induced

plasma polymerisation of propenoic acid and propanoic acid, J. Chem. Soc., Faraday Trans., 91
D

(1995) 3907-3912.
TE

[216] C.-H. Tsai, W.-J. Lee, C.-Y. Chen, P.-J. Tsai, G.-C. Fang, M. Shih, Difference in
P

Conversions Between Dimethyl Sulfide and Methanethiol in a Cold Plasma Environment, Plasma
CE

Chem. Plasma Process., 23 (2003) 141-157.


AC

[217] M. Alexander, T. Duc, The chemistry of deposits formed from acrylic acid plasmas, J.

Mater. Chem., 8 (1998) 937-943.

[218] B. Mitu, V. Satulu, G. Dinescu, Mass Spectrometry Diagnostic During RF Plasma

Polymerization of Thiophene Vapors, Romanian J. Phys., 56 (2011) 120-125.

[219] D. Thiry, F.J. Aparicio, P. Laha, H. Terryn, R. Snyders, Surface temperature: A key

parameter to control the propanethiol plasma polymer chemistry, J. Vac. Sci. Technol., A, 32

(2014) 050602.

109
ACCEPTED MANUSCRIPT

[220] S. Ershov, F. Khelifa, V. Lemaur, J. Cornil, D. Cossement, Y. Habibi, P. Dubois, R.

Snyders, Free Radical Generation and Concentration in a Plasma Polymer: The Effect of

Aromaticity, ACS Applied Materials & Interfaces, 6 (2014) 12395-12405.

PT
[221] P. Marsal, M. Roche, P. Tordo, P. De Sainte Claire, Thermal stability of OH and O-alkyl

RI
bonds in N-alkoxyamines. A density functional theory approach, J. Phys. Chem. A, 103 (1999)

2899-2905.

SC
[222] D. Barton, A.G. Shard, R.D. Short, J.W. Bradley, The effect of positive ion energy on

NU
plasma polymerization: a comparison between acrylic and propionic acids, J. Phys. Chem. B, 109

(2005) 3207-3211.
MA
[223] M.R. Alexander, F.R. Jones, R.D. Short, Mass spectral investigation of the radio-frequency

plasma deposition of Hexamethyldisiloxane, J. Phys. Chem. B, 101 (1997) 3614-3619.


D

[224] M. Alexander, F. Jones, R. Short, Radio-frequency hexamethyldisiloxane plasma


TE

deposition: a comparison of plasma-and deposit-chemistry, Plasmas Polym., 2 (1997) 277-300.


P

[225] S. Saboohi, M. Jasieniak, B.R. Coad, H.J. Griesser, R.D. Short, A. Michelmore,
CE

Comparison of Plasma Polymerization under Collisional and Collision-less Pressure Regimes, J.


AC

Phys. Chem. B, DOI 10.1021/acs.jpcb.5b07309(2015).

[226] J. Benedikt, Plasma-chemical reactions: low pressure acetylene plasmas, J. Phys. D: Appl.

Phys., 43 (2010) 043001.

[227] A. Howling, L. Sansonnens, J.L. Dorier, C. Hollenstein, Time‐resolved measurements of

highly polymerized negative ions in radio frequency silane plasma deposition experiments, J.

Appl. Phys., 75 (1994) 1340-1353.

[228] I. Swindells, S.A. Voronin, P.M. Bryant, M.R. Alexander, J.W. Bradley, Temporal

Evolution of an Electron-Free Afterglow in the Pulsed Plasma Polymerisation of Acrylic Acid, J.

Phys. Chem. B, 112 (2008) 3938-3947.


110
ACCEPTED MANUSCRIPT

[229] I. Swindells, S.A. Voronin, C. Fotea, M.R. Alexander, J.W. Bradley, Detection of negative

molecular ions in acrylic acid plasma: some implications for polymerization mechanisms, J.

Phys. Chem. B, 111 (2007) 8720-8722.

PT
[230] D. Barton, R.D. Short, S. Fraser, J.W. Bradley, The effect of ion energy upon plasma

RI
polymerization deposition rate for acrylic acid, Chem. Commun., (2003) 348-349.

[231] T. Cleland, D. Hess, In situ FTIR diagnostics of the radio-frequency plasma decomposition

SC
of N2O, Plasma Chem. Plasma Process., 7 (1987) 379-394.

NU
[232] J. Röpcke, G. Lombardi, A. Rousseau, P. Davies, Application of mid-infrared tuneable

diode laser absorption spectroscopy to plasma diagnostics: a review, Plasma Sources Sci.
MA
Technol., 15 (2006) S148.

[233] J. Nishizawa, N. Hayasaka, In situ observation of plasmas for dry etching by IR


D

spectroscopy and probe methods, Thin Solid Films, 92 (1982) 189-198.


TE

[234] C. Mogab, A. Adams, D.L. Flamm, Plasma etching of Si and SiO2—The effect of oxygen
P

additions to CF4 plasmas, J. Appl. Phys., 49 (1978) 3796-3803.


CE

[235] K. Takahashi, A. Itoh, T. Nakamura, K. Tachibana, Radical kinetics for polymer film
AC

deposition in fluorocarbon (C4 F8, C3 F6 and C5 F8) plasmas, Thin Solid Films, 374 (2000) 303-

310.

[236] T. Shirafuji, A. Tsuchino, T. Nakamura, K. Tachibana, Plasma copolymerization of

C6F6/C5F8 for application of low-dielectric-constant fluorinated amorphous carbon films and its

gas-phase diagnostics using in situ Fourier transform infrared spectroscopy, Japanese Journal of

Applied Physics, 43 (2004) 2697.

111

Potrebbero piacerti anche