You are on page 1of 42

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/299593562

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


critical review

Article in Thin Solid Films · May 2016


DOI: 10.1016/j.tsf.2016.02.058

CITATIONS
READS
58
2,227

4 authors:

Damien Thiry
Stephanos Konstantinidis
Université de Mons
Université de Mons
44 PUBLICATIONS 538 CITATIONS
91 PUBLICATIONS 2,294 CITATIONS

Jérôme Cornil
Rony Snyders
Université de Mons
Université de Mons
408 PUBLICATIONS 27,805 CITATIONS
176 PUBLICATIONS 2,594 CITATIONS

Some of the authors of this publication are also working on these related projects:

Pinning and Polarization Effects in Molecular Junctions View project

Physical Chemistry of Plasma-Surface Interaction - PSI View project


All content following this page was uploaded by Damien Thiry on 20 June 2018.

The user has requested enhancement of the downloaded file.


Thin Solid Films 606 (2016) 19–44

Contents lists available at ScienceDirect

Thin Solid Films

j our na l hom ep age : w ww . e l se v i e r . c o m / l o c a t e / t sf

Critical review

Plasma diagnostics for the low-pressure plasma polymerization


process: A critical review
Damien Thiry a,b, Stephanos Konstantinidis a, Jérô me Cornil c, Rony Snyders a,d
a
Chimie des Interactions Plasma-Surface, CIRMAP, University of Mons, Place du Parc 20, B-7000 Mons, Belgium
b
Institut des Matériaux Jean Rouxel, Université de Nantes, CNRS, 2 rue de la Houssinière B.P. 32229, 44322 Nantes Cedex 3, France
c
Service de Chimie des Matériaux Nouveaux (CMN), CIRMAP, University of Mons, 20 Place du Parc B-7000 Mons, Belgium
d
Materia Nova Research Center, Parc Initialis, Avenue Nicolas Copernic 1, B-7000 Mons, Belgium

a r t i c l e i n f o a b s t r a c t
Available online 28 February 2016 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 materials, controlling the chemistry of
Keywords: the coatings by a clever choice of the process parameters represents the main challenge. And yet, it became
Plasma polymerization
quick- ly obvious that in view of the complexity of the growth mechanism, fine control of the layers properties
Plasma polymer films
can only be reached by understanding at a fundamental level the mechanistic formation of the layers. In this
Plasma diagnostic
Growth mechanism
context, a de- tailed comprehensive study of plasma chemistry is therefore of crucial importance as the
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 polymerization process, namely
Mass Spectrometry, in- situ Fourier Transform Infrared Spectroscopy, Optical Emission Spectroscopy,
Langmuir and Ionic probes are reviewed. After a light description of each technique, the main achievements
for improving the mechanistic un- derstanding 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 highlighted. In view of the better control of the process allowed by the plasma phase
investigation, some general conclusions and perspectives de- scribing future developments in the field of
plasma polymerization are finally discussed.
© 2016 Elsevier B.V. All rights reserved.

Contents

1. Introduction.......................................................................................................................................................................................................................................... 20
2. Plasma polymerization......................................................................................................................................................................................................................... 21
2.1. Plasma fundamentals............................................................................................................................................................................................................... 21
2.2. Growth mechanism.................................................................................................................................................................................................................21
2.3. Control of PPF properties through deposition parameters.................................................................................................................................................... 23
3. Which tools can we use to probe the plasma polymerization process?.......................................................................................................................................25
3.1. Diagnostic techniques............................................................................................................................................................................................................25
3.2. Theoretical support for a better description of the data....................................................................................................................................................25
4. Overview of the diagnostics methods...............................................................................................................................................................................................26
4.1. Electrostatic probe................................................................................................................................................................................................................... 26
4.2. Optical emission spectroscopy..............................................................................................................................................................................................27
4.3. Mass spectrometry................................................................................................................................................................................................................... 30
4.3.1. Neutral species analysis............................................................................................................................................................................................ 30
4.3.2. Ion analysis................................................................................................................................................................................................................ 33
4.4. Ion probes................................................................................................................................................................................................................................. 36
4.5. Gas-phase Fourier transform spectroscopy............................................................................................................................................................................ 37
5. Conclusions and perspectives............................................................................................................................................................................................................40
Acknowledgments........................................................................................................................................................................................................................................41
References...................................................................................................................................................................................................................................................... 41

E-mail address: damien.thiry@umons.ac.be (D. Thiry).

http://dx.doi.org/10.1016/j.tsf.2016.02.058
0040-6090/© 2016 Elsevier B.V. All rights reserved.
D. Thiry et al. / Thin Solid Films 606 (2016) 19–44
2
1.Introduction Fig. 1. Schematic comparison of a plasma polymer film and a conventional polymer
material obtained from the same precursor/monomer.

The interactions of a solid with its surrounding are mainly defined


by the physico-chemical properties of its surface. It is therefore not
surpris- ing that since many decades, continuous research and
developments in materials science have contributed to the rapid
growth of surface and coating technologies which nowadays still
increasingly attract consider- able attention. Surface technologies refer
to the modification of the sur- face (e.g., chemical functionalization,
etching, coating,…) of a material without changing its bulk properties.
For instance, tailor-made coatings allow adjusting mechanical (wear,
friction), chemical (corrosion, per- meation, temperature insulation,
biocompatibility, wettability), electri- cal (conductivity), and optical
(transmission, reflection, absorption, color) properties of materials
[1].
Through the years, numerous processes have been developed for
the modification of surfaces via the synthesis of thin films. A non-
exhaustive list includes chemical vapor deposition, pulsed laser
deposi- tion, spin coating, sol gel, spin casting, thermal evaporation
and plasma- based technologies [1–4]. Among them, the plasma-based
processes are of particular interest by combining significant
advantages such as their low process temperature, enabling the
treatment of a wide range of ma- terials 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].
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 fabrica-
tion of antibacterial surfaces [12], protein biochips [13,14] or
platforms for biomolecules immobilization [15,16]. These surfaces can
also be syn- thesized using plasma-based technology, more
specifically by means of the plasma polymerization method, allowing
the formation of solid or- ganic thin films referred to as plasma
polymer films (PPF). Despite the use of the word “polymer”, PPF
present little resemblance to the con- ventional polymers except for
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 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 from which the material is
built. Nevertheless, both terms are accepted and currently employed
in the plasma polymerization community.
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.
D. Thiry et al. / Thin Solid Films 606 (2016) 19–44
Nevertheless, they were considered as a nuisance until the transform infrared spectroscopy and optical emission spectroscopy.
2
work of Goodman who demonstrated that a 1 μm thick plasma- The main results obtained by Langmuir and ionic probes to determine
polymerized styrene film deposited on a titanium foil made a the plasma parameters and ion flux, respectively, are also discussed.
satisfactory dielectric for a nuclear battery [19]. Since that time, In addition, throughout the paper, it is shown how theo- retical
the potential of these organic coatings has been revealed and a calculations based upon the density functional theory method have
systematic investigation of the plasma polymerization process has proven to be a powerful tool for assisting in the interpretation of
been carried out. More information about the history of the
plasma polymer science can be found in Refs. [20–22]. It is now
demonstrated and accepted that PPF exhibit interesting physico-
chemical properties for organic materials such as high thermal,
mechanical and chemical stabilities [4]. Moreover, due to their
intrinsi- cally good adhesion properties, numerous materials (e.g.,
glass, poly- mers, metals), even with complex geometry (e.g.,
carbon nanotubes [23–26], micro/nanoparticles [27–31]), 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 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 oc-
curs in the plasma. Thus, highly energetic conditions have usually
been employed due to the precursor fragmentation dependence on
the ener- gy dissipated in the system [21]. As a result of these
extensive fragmen-
tation reactions, poor control of the PPF chemistry was achieved.
Since the 1980s, with the rise of micro- and nano-technologies,
plas- ma polymerization has been further developed in the search
for PPF with controlled and tailored chemistry while keeping their
other inher- ent 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 this class of materials arises from their potential use in modern
fields of applications including the development of supports for
biomolecules immobilization [71–77] or cell growth [78–80],
interlayers for promot- ing adhesion of metal coatings [81],
biocompatible [82] or antibacterial coatings [12,83], controlled
drug release coatings [84–87], super- hydrophobic surface [62],
conductive layers [70],etc.
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 control of the plasma polymer chemistry. It quickly
appeared that this control can be obtained through the
understanding of the PPF growth mechanism and more
specifically of the phenomena taking place at the plasma–
substrate in- terface. Accordingly, investigating the plasma
chemistry rapidly turned out to be a necessity. Numerous works
have therefore been focused on the investigation of the plasma
phase during the PPF growth. Surpris- ingly, while other aspects of
the 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 de- scribe how these works have contributed
to enlarging the knowledge of plasma polymer growth at a
molecular level.
This review is organized as follows. In the first part, the plasma
poly- merization 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 most popular diagnostics methods
employed for probing the spe- cies constituting the plasma are
described, namely, mass spectrometry, gas phase Fourier
the complex diagnostic data. Finally, conclusions and perspectives
sug- gesting research strategies for increasing the fundamental
knowledge in the plasma polymer field are given.

2.Plasma polymerization

In this section, we start with a brief summary of the main features


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


Fig. 2. Typical Maxwell–Boltzmann electron energy distribution function for low pressure
discharges. Adapted from [10,20].
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
in Fig. 2 [20,92]. Although in a conventional plasma polymerization
first described in 1920 by Irving Langmuir working on the
pro- cess, the majority of electrons possesses a kinetic energy
development of vacuum tubes for large currents [11]. Although
omnipresent in the universe and representing nearly 99% of the centered around a given value (1–2 eV in Fig. 2), the distribution
extends to high energy (from 4 to 20 eV). This allows the occurrence
matter (solar corona, solar wind, earth's ionosphere), the natural
of numerous chemical reactions in the plasma through collisional
presence of plasma on earth is rare (e.g., lightning, aurora borealis)
processes with electrons resulting in the formation of a great variety
[93]. Nevertheless, the plasmas can be easily generated artificially via
of species includ- ing radicals, ions and photons. Since the degree of
the electrical excitation of a gas, for example.
ionization in the plas- ma is often very low, the most probable
The particles (electrons, ions and neutrals) constituting the
collisions involving electrons occur with neutrals particles. The
plasma are in motion and undergo collisions during which energy is
probability for a given reaction to occur depends on the number of
transferred. Each family of particle is often characterized by a
electrons able (from an energetic point of view) to induce the
temperature related to its translational energy. In plasma science, the
chemical reaction. On this basis, consider- ing organic discharges, in
temperature of the parti- cles is generally expressed using the eV as
comparison to ionization, dissociations reac- tions leading to the
unit; the conversion factor is 1 eV = 11600 K. Depending upon
formation of radicals are more probable than ionization events since
potential differences in terms of temperature for the particles
most of the electrons in the plasma possess an energy of 1–4 eV
constituting the plasma, they are classified in two main families:
similar to the energy required to break simple organic chemical bonds
namely, at thermodynamic and at non- thermodynamic equilibrium.
(see Fig. 2); in contrast, the ionization energy of organic molecules is
For the former case, all the particles are characterized by a unique
approximately 10 eV. As a consequence, the concentration of radicals
temperature, namely Te = Tn = Ti where Te, Ti and Tn represent the
in the plasma can be 103–105 higher than the ion density [22].
temperatures associated to electrons, ions and neutrals species,
Another important aspect related to plasma technology is that
respectively. In contrast, for non-equilibrium
owing to the significant difference in terms of thermal flow velocities
plasmas, the electron temperature strongly differs from that
associated with the heavy particles, ions and neutrals: T e ⋙ Ti ≥ Tn. between electrons and ions, a surface immersed in a plasma naturally
acquires a negative charge which acts to accelerate the positive ions
Indeed, where- as the electron temperature typically ranges from 1 to
10 eV, the ion and toward the surface and repel most of the impinging electrons [8,92].
neutral temperatures are close to room temperature, 0.025 eV (298 For insulated surfaces (as generally encountered for plasma polymer
K). This is why such plasmas are often called “cold” plasmas. Indeed, layers), at the equilibrium, the ion flow perfectly counterbalance the
although the temperature of the electrons is high, their low density electron flow leading to a stationary electrical potential named the
and heat capacity allow surfaces surrounding the plasma to remain at floating potential (Vf). As a consequence of this phenomenon, near all
relatively low temperatures [11]. This non-equilibrium feature is surfaces including the substrate, regions called sheath characterized
partic- ularly attractive for materials processing as the electrons can by a net positive charge develop. When ions enter the sheath, they are
induce several chemical reactions without altering the treated accelerated and strike the surface with a kinetic energy ranging
material by excessive temperature. Consequently, cold plasmas are typically from 10 to 30 eV in the case of a floating surface in the
predomi- nantly used in surface material processing including plasma absence of collisions, i.e., when considering a low pressure plasma
polymerization [91]. For conventional plasma sources (i.e., direct polymerization process [38]. The influence of the bombarding ions on
current, radio-frequency, microwave), the ionization degree is less plasma polymer- ization is obviously also crucial and is discussed
than a few percent. Cold plasmas can be generated either at low below. It should be noted that when the substrate is directly connected
(i.e., b 1 Torr) or high pressure, even at atmospheric pressure. In the to an RF power sup- ply, the negative self-bias developed at the
context of the plasma polymerization process, most studies are con- surface can be significantly higher than the floating potential and
ducted at low pressure and are the subject of this review. However, depends on the synthesis condi- tions [8]. Finally, another source of
there is recently an increasing interest for operating the plasma poly- energy is also provided to the grow- ing film due to the irradiation of
merization process at atmospheric pressure. More details about this ultraviolet (UV) photons emitted by excited atoms/molecules.
aspect can be found in the review of Merche et al. [94].
Since the electrons are the energy vectors in the plasma, their 2.2. Growth mechanism
density and energy are of crucial importance for processes occurring
in the gas phase. The electrons are assumed to be in thermal Their unique properties make the PPF a specific class of materials
equilibrium at a given temperature Te [92]. Following kinetic gas arousing, over many decades, the curiosity of surface scientists who
theory, the electrons energy distribution function (EEDF) can be, to a have developed strong efforts to understand their growth mechanism.
first approxi- mation, expressed by a Maxwell-Boltzmann distribution
as illustrated
The overall plasma polymerization mechanism involves both Fig. 3. Schematic description of the rapid step-growth polymerization mechanism. Adapted
from [96].
gas phase and surface reactions [95]. The first step consists in the
vapor- ization of an organic precursor in a deposition chamber. 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 precursor
dissociations reactions resulting in the formation of radicals.
Historically, it was therefore as- sumed that the growth of the layer
mainly occurs through either radical-radical or radical-molecule
reactions. Indeed, the radicals pos- sess an unpaired electron and
therefore are highly reactive toward ter- mination reactions with
other radicals or toward addition reactions with 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
func- tion, numerous fragmentation pathways are possible [66]. As a
conse- quence, a great variety of radicals is generally produced in the
plasma whatever the chemical nature of the organic precursor, hence
contribut- ing to the difficulty to obtain fine control of the PPF
chemistry.
Some of the processes involved in the synthesis of a PPF are
summa- rized in the “Rapid Step Growth Polymerization” (RSGP)
model pro- posed by Yasuda in 1985. The model is built on the
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 where- as cycle 2 concerns biradical species, both
originating from the interac- tion of the precursor with the plasma.
Steps (1), (3), (4) and
(5) correspond to addition 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 reactive species including radicals and biradicals
(steps (3) and (5)). In both cases, the products formed can undergo
other propagation reactions. It is obvious that the extent of addition
reactions in the gas phase depends on the pressure.
In the plasma polymerization mechanism, the molecules formed
through the recombination between two radicals (step (2),
termination reaction) can be reactivated via electron impact in
contrast with the sit- uation encountered in conventional
polymerization. Consequently, the RSGP mechanism can be viewed as
a succession of termination reac- tions followed by the reactivation of
the products. When a surface is ex- posed to such a plasma, a solid
organic thin film arising from the condensation of the reactive species
produced in the plasma is deposit- ed on the substrate. However,
further “polymerization” reactions (initi- ation, addition, termination,
reactivation) can still occur at the plasma growing-film interface.
In order to provide a more complete description of the plasma
poly- merization mechanism, Yasuda added in his model the CAP
(“Competi- tive ablation and polymerization”) principle, namely the
simultaneous occurrence of film etching and deposition processes
[21,39,96]. The syn- thesis of the coating thus results from a
balance between both
phenomena. Indeed, highly reactive radicals (e.g., .O., .F, .S., etc) can synergetic effect between ions and neutrals has to be considered since
be produced and react at the interface to form stable molecules increasing the ion flux en- ables the formation of a higher number of
(e.g., CO2, CO, CS2, CF4). The products may desorb from the growing reactive sites at the surface, thus promoting the grafting of radicals.
film leading to the removal of part of the coating (ablation). These To summarize, although nowadays it is accepted that the contribu-
stable molecules cannot take part in the growth of the film tion of ions in the deposited film is significant, their exact role in the
anymore and are either
(i) pumped out of the reactor or (ii) reactivated through electron
colli- sions. The occurrence of such surface reactions directly
depends on the amount of energy provided to the growing film
interface by ion bom- bardment and UV photon irradiation. It
should be noticed that these sta- ble molecules can also be formed
via gas phase reactions reducing 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 compared to radicals and neutrals, the
influence of ions in the growth mechanism has been totally
excluded. Nevertheless, as al- ready mentioned, in plasma
processing, a surface exposed to the dis- charge 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 breaking lead- ing to the
formation of surface dangling bonds which can act as prefer- ential
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
the formation of surface defects through ion interaction is
considered [60,98]. Based on the 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 plasma polymeri- zation 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 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 com- pleteness, 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 de- posited by ions would be insignificant [100]. However,
experimental ev- idence has clearly revealed that the ions play a
greater role in the 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 [57]. The
origin of the greater role played by the ions in PPF formation is
justified by several factors. For instance, the presence of a sheath
at the plasma growing film interface accelerates the ions toward
the film sur- face, significantly reducing the neutral-to-ion flux
ratio at the surface in comparison with that measured in the gas
phase [8,102]. Considering a low pressure plasma, e.g., 1 mTorr,
the surface flux ratio of neutrals to ions is estimated to be
approximately 20 times lower than the corre- sponding density
ratio in the plasma [102]. Another important factor to take 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
depending on their en- ergy [105,106]. On the other hand, for
radical-based molecules, the
sticking coefficient has been estimated to vary over a wide
range (~ 10−4 to ~ 1) depending on the chemical nature of the
radical (e.g., unsaturation degree) [ 1 0 7 – 1 0 9 ] and the activation
of the surface
through the formation of open bond sites at the interface
[99,107,110, 111]. Therefore, as integrated in the AGM, a
plasma polymerization mechanism is still debated [112,113]. It is,
parameters and PPF properties. Very often, an empirical approach is
how- ever, important to shed light on their impact for defining PPF
uti- lized and the common strategy consists in separately studying the
properties as recently demonstrated by Michelmore et al. for the
influ- ence of one process parameter, while keeping the other ones
mechanical prop- erties of the deposited layers [58].
constant. For a given reactor configuration, the influence of the
The combination of the different models described in this section
operational pa- rameters can be understood by considering their
(schematically represented in Fig. 4) allows us to provide an overview
impact on the plasma characteristics. For example, the applied power
of the main reactions taking place in the plasma and at the PPF
alters the electron ener- gy distribution function and density which
surface. This complex mechanism is at the origin of the poor control
govern the nature and the rate of the chemicals reactions occurring
of the chemistry of such thin films. Indeed, it has been extensively
in the plasma [142]. This, in turn, influences the chemistry of the
reported that functionalized PPFs contain numerous chemical
coatings and their physico-
functions even if the organic precursor is monofunctional [21,67].
chemical properties.
This complex mecha- nism also explains the high branched and cross-
According to the plasma polymerization mechanisms, the most im-
linking degree of plas- ma polymers triggering their attractive
portant factor affecting the properties of a PPF is the energy applied
physico-chemical properties. This irregular structure results from the
per molecule which governs the degree of precursor fragmentation in
random recombination of the numerous plasma-generated fragments
the discharge [96]. In this context, Yasuda proposed a composite
followed by successive rear- rangements, fragmentations,
param- eter (the so-called Yasuda parameter): W/FM where W, F and
reinitiations reactions, etc.
M are the power dissipated in the discharge in J/s, the monomer flow
Another consequence of this complex mechanism is the presence
of rate in mol/s and the molecular weight of the monomer in kg/mol,
respectively [97]. The Yasuda parameter represents the energy input
radicals in the PPF network after synthesis [72,114–118]. These
per unit mass of monomer. The term W directly scales with the
radicals have been identified to be at the origin of the well-known
electron density and hence governs the collision frequency between
“aging” of PPFs, namely oxidation of the material [49,82,114,119].
electrons and precursor molecules [40,95]. For a fixed working
Nevertheless, it has been recently reported that they can be
pressure, the term 1/F scales with the residence time of the particles in
advantageously exploited for inducing a radical conventional
the discharge, thus also influencing the extent of precursor
polymerization [117,120–124]. The trapped radicals can also serve as
fragmentation. Therefore, the concentration of activated species in
reactive sites for the grafting of proteins [125,126]. Another aging
the plasma directly depends on the W/FM parame- ter. Regarding the
phenomenon is related to the potential presence of embedded
evolution of the deposition rate with W/FM, two prin- cipal domains of
molecules in the plasma polymer matrix [49,82,127] which can be
plasma polymerization can be identified (see Fig. 5). At low W/FM, the
released in air during storage of the material [66,128–132] or after
deposition rate evolves linearly with W/FM. In this re- gion, the so-
immersion of the coatings in a liquid medium [36,49,68,133–135].
called “energy-deficient domain”, ample monomer is avail- able and
The latter could be detrimental for biological applications requiring
the supply of energy is the limiting factor for increasing the
the immersion of the layers in solution [82,127,136,137].
deposition rate. In this domain, increasing the supply of energy in the
discharge results in an increase in the concentration of film-forming
2.3. Control of PPF properties through deposition parameters
species through collisional processes. Above a critical W/FM value,
the deposition rate becomes constant since the precursor
The chemical composition of the layers (e.g., the density of a par-
fragmentation has attained its maximum. This regime is called
ticular chemical function) as well as their cross-linking degree are
“monomer-deficient” as sufficient energy is provided, but the
strongly affected by the deposition parameters: the plasma source
precursor feed rate into the chamber is the limiting factor [143]. In
(e.g., radiofrequency, microwave,…), the power dissipated in the
addition, under high energy con- ditions, ion-induced etching reactions
discharge, the working pressure, the precursor flow rate, the sub-
can be favored and this leads to a decrease in the deposition rate. The
strate temperature, etc. [4,8,17,21,22,131,138–140]. In addition, extent of ablation phenomenon strongly depends on the chemical
some geometric parameters related to the design of the chamber also composition of the discharge. It is, for example, well known that
affect the coating properties. Examples include the location of the pre- fluorine-based discharges are quite sensi- tive to ablation reactions
cursor inlet, the distance between the plasma source and the [4,61].
substrate, etc. The dependence of the process with the reactor
Complementary to the Yasuda concept and as a further develop-
geometry makes therefore difficult to compare the “same” PPF from
ment, Hegemann et al. developed a macroscopic approach for describ-
one deposition chamber to another [22,141].
ing the plasma polymerization process based on the concept of
The need for PPFs with a precise chemical composition and
chemical quasi-equilibrium [39,40,95,144–147]. In this case, the
structure requires a clear understanding of the relationship between
plasma is divided into (i) an active zone in which the activation takes
the process
place through collisional processes leading to the production of film-
forming species and (ii) a passive zone where the deposit is formed.
This approach states that the mass deposition rate Rm (expressed in

Fig. 5. Typical evolution of the deposition rate as a function of the Yasuda parameter,
Fig. 4. Overall mechanism of plasma polymerization.
W/FM illustrating the different regimes of plasma polymerization [143].
g/cm2s) depends on the parameter W/F following a quasi-Arrhenius methane discharge. Adapted from [146].

be- havior according to Eq. (1) [144]:

Rm Ea
¼ G exp − ; ð1Þ
F W=F

where G is a reactor and process dependent factor related to the


maxi- mum monomer conversion into film growth and Ea represents
an apparent activation energy related to the PPF system. It should be
em- phasized that the W/F parameter can be adjusted depending on
the re- actor 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
in- verse specific energy, (W/F)− 1), is shown in Fig. 6 for the plasma
poly- merization of methane. Ln (Rm/F) evolves linearly with (W/F)
−1

pointing to an Arrhenius-like behavior over the range of W/F


investigat- ed. In this regime, the film grows mainly through radical
reactions and the kinetic limiting factor is the production rate of the
film-forming spe- cies 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) holds for
many monomers and gas mixtures for a certain range of W/F
[146,147,150]. In Fig. 6, from the negative slope of the linear fit, Ea,
which is related to an ensemble of fragmenta- tion reactions in the
active zone, can be deduced [151]. Its value is monomer specific and
correlates with the bond dissociation energy of the precursor [146].
Compared to the Yasuda approach, the activation energy separates
the plasma polymerization into the energy and mono- mer deficient
regime [144,148].
When applied to several PPF families for a broad range of energy
conditions, the macroscopic approach has revealed that more than
two deposition regimes (i.e., the energy and monomer deficient
regime at high and low energy levels, respectively) have to be
considered de- pending on the plasma polymerization system. This
could be related to different growth 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 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 measurements. This has led to a very
in- teresting debate in the plasma polymer community regarding the
suit- ability of the macroscopic approach [39,148,151–154].

Fig. 6. Arrhenius-like plot of the deposition rate versus the inverse energy input for a
The energy applied to the plasma polymerization process a power equivalent to the power applied during the plasma “ON” time
affects not only the deposition rate, but also the composition and [37]. During this period, the layer grows through the complex
the cross-linking degree of the layers. At high power, more cross- mechanism previously
linked PPFs are formed because of the extensive fragmentation of
the precursor yielding a higher 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 growing film (see below).
These high energy conditions are suitable for obtaining good
barrier properties finding applications in food packaging and cor-
rosion protection [155]. On the other hand, plasma polymerization
con- ducted at low energy conditions provides a low degree of
precursor fragmentation and a high retention of the functional
group hosted by the precursor. This can be explained by the small
amount of electron- induced collisions with precursor molecules
leading therefore to an ac- tivation process rather than a complete
disintegration of the initial chemical structure. This mode of
operation is therefore more suitable for biological applications
[8,21,41,47,48,55,72,82,156].
As already mentioned, the energetic conditions at the growing film
surface due to ionic bombardment are a key factor controlling film
prop- erties. Therefore, in complement to the macroscopic
approach, with the aim to rationalize the influence of the ion
bombardment in plasma po- lymerization 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 following Eq. (2) [157]:

pffiffiffi ffiffiffiffiffiffi Γipffiffiffimffiffiffieffiffiaffiffinffiffi


π ¼ m ; ð2Þ surf i

R
where mi, Γi, 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 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 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 ener- getic 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 pro- cess 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 conditions, an
irregular structure predominates and the density of the targeted
function remains low. With this prob- lem in mind, a breakthrough
appeared when the pulsed plasma tech- nique, reported in the
field of plasma polymerization in 1972 by Tiller et al., was
introduced [21]. The idea was to further reduce the extent of
fragmentation of the precursor and hence the side reactions
accounting for the formation of other functionalities than the one
hosted by the precursor. Today, pulsed plasma polymerization has
be- come a well-established method for the synthesis of
functionalized PPF [37,41,42,44,46,48,54,59,68,70,159–161].
The pulsed approach consists in producing the discharge
intermit- tently according to the pulse 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].
During the plasma “ON” time, electrons, ions, radicals and
photons are produced. The ton period can be approximated to a
discharge oper- ating in Continuous Wave (CW) mode sustained at
discharge in the CW mode can be employed, thus allowing a
significant reduction in the degree of fragmentation of the precursor
and in turn enhancing the retention of the chemical group of interest
[37,46].

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 other and with
the grow- ing film via a multitude of interaction pathways [117]. This
complex mechanism makes the assessment of each individual
reaction as well as the prediction of the coating properties very
challenging. Further- more, depending on the chemical nature of the
precursor and the syn- thesis conditions, especially at low energy
conditions, some specific reactions can predominate. This would also
influence the final features of the formed layers. 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].

3.1. Diagnostic techniques

When concluding one of his papers dealing with the development


of the macroscopic approach, Hegemann wrote: “ It is expected that
further clarification of the complex processes taking place during plasma
polymer- ization can be achieved by the use of plasma diagnostics in
combination with the macroscopic approach” [146]. Therefore, in
addition to the de- velopment 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 spec- trometry, gas phase Fourier Transformed
Fig. 7. (a) Plasma polymer film (from C 2H4, NH3/C2H4 and CO2/C2H4) density vs. the
Infrared Spectroscopy, optical emission spectroscopy, electrostatic and
momentum density during film growth. The linear evolution indicates densification by
momentum transfer. (b) Chemical composition of a plasma polymer from NH 3/C2H4 and ion probes) and surface analy- sis (X-ray Photoelectron Spectroscopy-
CO2/C2H4 depending on the momentum density during film growth. The filled symbols XPS, Time of Flight Secondary Ion Mass Spectrometry-ToF-SIMS)
indicate the relative amount of oxygen and nitrogen. The open symbols show the tools.
functional group density which is reduced with the increasing densification. Adapted Several diagnostic methods have been developed and employed
from [157].
during the last 30 years for detailed comprehensive study of the
plasma chemistry. For evaluating the chemical composition of the
plasma, non- intrusive optical diagnostics methods including optical
described. When the plasma is switched OFF, due to the emission spec- troscopy (OES) and gas-phase Fourier Transform
recombination of the electrons and ions at the reactor walls, the Infrared Spectroscopy (GFTIR) as well as mass spectrometry (MS)
electron density as well as the floating potential naturally developed at have been used extensively. In complement to these techniques, ion
the substrate rapidly de- crease as experimentally measured [41]. probes, especially designed for operating in organic discharges, have
Consequently, in this case, ion bombardment and photon irradiation also been developed with the aim to measure absolute ion fluxes
rapidly disappear, limiting the side reactions at the surface and thus toward surfaces facing the plasma. Final- ly, governing the production
favoring the reaction of radicals present in the gas volume with the rate of species in the plasma, the electron density and temperature
nucleation sites generated during ton. can be evaluated using electrostatic probes. Table 1 summarizes the
Depending on the chemistry of the starting molecule, the main features of the diagnostic methods employed in the context of
precursor itself can participate in addition reactions with the radicals low-pressure plasma polymerization. The techniques are described in
sites present at the surface of the growing film even during toff. This more detail in the next section of this review.
has been exper- imentally demonstrated for the pulsed plasma
polymerization of acrylic acid [37]. This precursor, containing a 3.2. Theoretical support for a better description of the data
double bond, can therefore easily undergo a propagation reaction
through a radical mechanism involving the opening of the double If the evaluation of the plasma chemistry is a challenging task, the
bond. In this case, the pulsed mode can ideally be viewed as a in- terpretation of the accumulated data is often, at least, as difficult.
succession of activation reactions (ton) followed by prop- agations This is why the use of theoretical tools is relevant in this context when
steps during toff, thus favoring the incorporation of the targeted used in close synergy with experimental measurements [179].
chemical group in the coatings [45]. Modeling the plasma chemistry is a formidable task in view of the
If non-classical polymerizable molecules are considered, growth large number of species involved and the large diversity of intervening
events during toff can be excluded. Indeed, in this case, during t off, the processes (inter- action between light and plasma species, heat
radicals rapidly recombine at the interface and the precursor itself propagation, chemical re- actions, thermal diffusion of the plasma
does not participate in the formation of the film, as revealed for the species, etc.). Many theoretical studies rely on mechanistic models
pulsed plasma polymerization of 3-fluoroaniline [160]. Nevertheless, involving generally only a fraction of the processes and/or species,
even in this case, the use of the pulsed plasma polymerization without any atomistic detail; in practice, kinetic equations are solved
strategy is beneficial as mean powers lower than those required to to describe, as a function of time, for in- stance the generation of the
sustain the different species in the plasma (together with their degree of
charging) and their spatial distribution. The main limitation of such
approaches is that the results heavily depend on the
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, ++ = medium relevance tool, +++ = high relevance tool.

Method Species probed or measurable parameters Time-resolved Comments Relevance in view of the
elucidation of the growth
mechanism of PPF

Electrostatic or
Electron density/temperature, plasma/floating
Langmuir probe Yes [41,162,163] +
Potential [142,162]
Semi-quantitative in some cases if using
OES Excited neutrals/ionic species [69,70,132,156,164,165] Yes [166] ++
an internal standard [156,166–168]
–Neutrals and ions depending on the analysis –Semi-quantitative [48,54,66,68,131,172]
mode [37,38,54,57,58,66,68,101,128,169,170] Yes for ions
Mass spectrometry +++
[41,163] –Quantitative for neutrals in some particular
–Ion energy distribution function of ions [38,104,171] cases [173–175]
Ion probe Absolute ion flux toward a surface [10,101,102,104,176] No ++
GFTIR Vibration frequency of chemical bonds [52,99,177,178] No Quantitative using a calibration procedure [52,99] +++

actual values chosen for the kinetic rates, in particular for the of an additional CH 2 group on the surface starting from a CH 3 radical
underlying activation energies. Those rates can be inferred from [182]; in this study, DFT was actually coupled with a force field
experimental mea- surements or estimated from sophisticated (within a Quantum Mechanics/Molecular Mechanics—QM/MM—
quantum-chemical calcula- tions. A critical comparison between approach) in order to treat the bulk diamond material at a lower level
reference data and the results of the simulations is required to of theory to ac- count for the electrostatic environment in the
validate the chosen parameters. A refine- ment is to account for the simulations and the part of the surface where the grafting occurs with
chemical structures by using molecular dynamics simulations based DFT to access electronic properties. Another study reported the use of
on force fields (i.e., expressions yielding the relative energies of a DFT calculations to select the best precursors for the plasma
given system in different geometries based on a series of bonded and polymerization of organosilicates by computing bond dissociation
non-bonded energetic terms). These methods cannot account for energies and free enthalpies of reac- tions [183]. Interestingly,
electronic excitations of the species or for the pres- ence of free algorithms to find transition states (and hence to estimate activation
electrons; moreover, although standard force fields cannot describe energies) are now implemented in many quantum-chemical packages,
the formation of chemical bonds, the implementation of reac- tive using, for example the intrinsic coordinate reaction (IRC) theory
force fields (such as ReaxFF force fields [180]) allows one to circum- [184]. The transition state can be validated by:
vent this limitation. Such simulations are deterministic in the sense (i) a frequency analysis showing a negative frequency for one mode,
that they can describe the trajectory of all species. There are, and (ii) by a steepest descent algorithm from the transition state
however, two main limitations associated with such simulations: (i) ensur- ing that the correct reactants and products are reached. This
the quality of the results critically depends on the nature of the approach has been exploited in a study modeling the production of
interatomic potentials used to describe the van der Waals interactions ethane from methane in a plasma [185]. Interestingly, all these
and the charge assign- ment on the atoms to evaluate the Coulomb studies involved the B3LYP functional of DFT owing to its good
energies; (ii) those simula- tions cannot be solved analytically and performance in reproducing experimental enthalpies of formation
require a time discretization, with a very small time step (of the order [186].
of a femtosecond). Accordingly, the simulations are typically run over Based on the previous considerations, the DFT method appears to
a limited time (typically a few hundred nanoseconds) and cannot be promising for a detailed description of the chemical reactions
account for slow kinetic processes. However, exploration of the taking place in the plasma. For a decade, our group has developed
conformational space can be accelerated by increasing the several strat- egies using DFT calculations in combination with
temperature or by coupling the force field to Monte Carlo simulations standard diagnostic techniques to enhance the understanding of
based on a random (compared to thermodynamic) sampling of the plasma chemistry, as it will be exemplified in the next section.
system.
A theoretical approach much less adopted in the field of plasma
4.Overview of the diagnostics methods
chemistry is to perform quantum-chemical calculations to shed light
on important properties of the plasma chemistry. The most popular
In this section, each diagnostic technique is individually presented
method used is Density Functional Theory (DFT) which limits the size
discussing its advantages and drawbacks. After a brief description of
of the system to be considered (typically up to one hundred heavy
the basic principle of the method, particular attention is devoted to
atoms), but provides key information such as the free enthalpy of
the link between diagnostic data and growth mechanisms. Further-
chem- ical reactions or activation energies. DFT is not a universal
more, in several examples, we also show how the use of DFT calcula-
theory and actually comes with many different flavors based on the
tions can be employed for assisting and completing the complex
choice of the functional and basis set. It is thus always
diagnostic data.
recommended to validate the selected DFT approach by comparison
to experimental data or highly sophisticated Hartree–Fock based
calculations. DFT can also be coupled to molecular dynamics 4.1. Electrostatic probe
simulations, for instance using the Car–Parrinello method though at a
much more extensive computational cost compared to force-field The electrons are the energy vehicles in the plasma and thus
calculations [181]. DFT typically sheds light on a very specific aspect in govern the production rate of reactive species through the
the plasma, generally to assess the change in the electronic energy (or dissociation/ioniza- tion of the precursor. Therefore, the knowledge
in the enthalpy or free enthalpy) associated with a reaction and of the electron temper- ature, density and electron energy
activation energies related to computed reaction pathways. For distribution function (EEDF) are mandatory for clarifying chemical
illustration, DFT has been used to model the growth of CVD diamond, reactions pathways. These plasma parameters, together with the
showing the different steps allowing for the insertion floating potential, can be obtained by
using the well-known electrostatic probe (also referred to as a Fig. 8. Evolution of the probe characteristic for Ar/H 2O and Ar/H2O/hexane gas mixtures
[188].
Langmuir probes).
This method is relatively easy to implement during PECVD experi-
ments. The probe itself 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 (regard-
less of their charge state or their chemical nature) and, as the voltage
be- comes positive, plasma electrons. By processing the current–
voltage characteristics, one can obtain plasma parameters as a
function of pro- cess parameters. More information on the theory as
well as technical as- pects of probe diagnostics can be found in Refs.
[92,187].
Langmuir probes have been used for a long time to analyze
deposi- tion plasmas whatever the process, i.e., Physical Vapor
Deposition or PECVD. However, in each case, scientists have to face
the same prob- lems, namely: the deposition of the coating onto the
probe surface and the probe body and the fact that the measurement
is spatially local- ized. The coating deposition issue 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 have devised
and upgraded their tools in order to minimize coating deposition
related issues, e.g., by programming cleaning routines between data
acquisition steps. The later can be achieved by positively biasing the
probe for several seconds.
Biederman et al. used a Langmuir probe during the DC deposition
of
hydrophilic films using hexane/Ar/H2O mixtures [188]. Fig. 8 shows
how the (absolute) probe current varies as a function of the probe
volt- age for various gas mixtures.
From their measurements, one can learn that the electron popula-
tion is divided into two groups, slow and fast electrons, when the
plas- ma is ignited in pure argon gas (24 Pa). The low 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 density (ne) varied from ~ 3 × 1015 m−3 to less than ~1 × 1015
m−3 as the per- centage of H2O was 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
dramatically as hexane was added to the mixture; the electron
temperature decreased to less than 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 the plasma. Obviously, these modifications
will change how the plasma spe- cies interact with the chamber walls
and the growing plasma polymer film.
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 well as the plasma spectroscopy (OES) which is based
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
compen- sated or uncompensated. Most of 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 har- monics, 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
temper- atures peak during the ON-time, when the electrical
energy is trans- ferred to the plasma species. The electron
temperature and density then quickly decay as the power supply
is switched off. As a conse- quence, the gas phase chemistry,
which is mainly promoted through electron impact, can be
modulated by varying the duty cycle (t on/ 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 in- creased. The
pulse amplitude and the duty cycle are thus key parame- ters when
devising the synthesis of plasma polymer films through PECVD
processes.
In order to minimize the probe coating deposition issue, a
modifica- tion 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 voltages 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 or- ganic plasmas [95]. More information on this
interferometry technique can be found in Ref. [189]. This method
offers some advantages as com- pared to the more conventional
electrostatic probes, namely (i) the mea- surement is not perturbed by
film deposition and (ii) it is not necessary to take into account the ion
composition in the plasma sheath in order to de- duce electron
density of the plasma. The main disadvantage lies in the lack of
information about the electron temperature.
Using microwave interferometry, Guimond et al. measured more
than a ten-fold increase in electron density, which increased from
~ 2.5 × 1015 m−3 to 3.5 × 1015 m−3 as the power was increased
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 deliv-
ered to the plasma reached ~ 60 W. Gas heating is invoked to
explain this behavior. Similar trends in the electron density vs
applied power were obtained by the same authors in C 2H4 and
C2H4/CO2 discharges [39]. These experimental results were used to
further describe the growth mechanism and to distinguish the
relative influence of both chemical and physical plasma processes.
In conclusion, Te, ne and EEDF are key data that should be known
in order to understand the fragmentation and excitation/ionization
reac- tions taking place in the plasma, and in fine, to better
understand the mechanistic formation of the film. Electrostatic
probes represent a sim- ple 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 imple- mented, such as the
microwave interferometry.

4.2. Optical emission spectroscopy

Investigating plasma chemistry is essential for identifying


reactive species taking part in PPF formation, and hence controlling
film proper- ties. Among other techniques (i.e., mass spectrometry
and the gas phase Fourier transform infrared spectroscopy detailed
in other sections), this can be achieved by optical emission
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 the duty cycle is set to 50%. Black circles, lozenges and triangles correspond to a peak power of 5, 20 and 40 W, respectively. Adapted from [162].

on the collection, by a spectrometer, of the radiation coming from the


the optical spectra becomes crowded by emission bands originating
plasma. Optical emission from a plasma occurs primarily through
from the multiple electronic, vibrational, and rotational excited states
elec- tron impact excitation of atoms or molecules according to Eq.
of the precursors. Fig. 10 shows a typical emission spectrum of ace-
(3):
tone/CO2 plasma.
N þ e− →Nω þ e− ; ð3Þ Although, numerous emission bands and lines overlap and are
poor- ly, or not, resolved, some lines can be unambiguously identified
where N* is the excited state of the species N. such as Hα, OH, CH, CO, CO2 and N. The presence of nitrogen emission
This is followed by a relaxation to a lower energy state releasing a might originate from impurities incorporated during the deposition
photon of energy equal to the difference between the two energy process. The presence of the other species result from the
states (Eq. (4)): dissociation/rear- rangement reactions taking place in the plasma,
hence highlighting the complex gas-phase chemistry in this kind of
Nω→N þ hν; ð4Þ plasma. Therefore, OES allows for a qualitative description of the
plasma chemistry en- abling the investigation of the chemical reaction
where h is Planck constant and ν is the frequency of the emitted pathways occurring in the discharge as reported for nitrogen
photon. The emission of specific frequencies can be used to identify the [23,192–195], fluorine [196,197], sulfur [69,70] and oxygen-based
spe- cies present in the studied plasma. More details about the [156,165–167,198–201] discharges.
technique can Another example that deserves to be pointed out is the detailed
be found in the following Refs. [92,187,190,191]. study of Granier et al. dealing with the investigation of plasma
OES studies are well documented in the literature related to low chemis- try by OES of hexamethyldisiloxane (HMDSO) and
pressure sputtering plasmas such as magnetron discharges because, tetraethoxysilane (TEOS) discharges with and without oxygen [199].
in this situation, the spectra are usually sparse and relatively easy to It has been shown that HMDSO plasmas are dominated by Si, SiO and
analyze. Indeed, most of the excited species inside such plasmas are SiH species whereas2 OH, CO, CO+ and CO+ emission lines are mainly
sputtered metal atoms and argon atoms. Diatomic molecules such as identified in TEOS plas- ma. From these data, the authors concluded
oxygen or nitrogen are added to the argon background gas if oxide or that the presence of CO and OH molecules in TEOS plasmas originates
metal nitride compounds have to be synthesized. In the case of from surface reactions at the plasma/growing film interface followed
plasma polymerization, which makes use of more complex organic by their desorption. These findings have allowed providing relevant
molecules, information on the growth mechanism of the coatings, an essential
step for understanding film properties. Nevertheless, it has to be
stressed that given the complexity of 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.
On the other hand, OES can also be used to monitor other important
features of the plasma, for instance the capacitive-to-inductive transi-
tion when working with inductively-coupled plasma 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 environ-
ment, the capacitive discharge is characterized by a weak global emis-
sion 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 in- tensity, ii) an increase in the PPF deposition rate and iii)
a decrease in the sulfur content in the PPF. Such behavior is related to
the intense fragmentation of the precursor when more energy is
dissipated in the plasma.
Fig. 10. Emission spectrum recorded in the ultraviolet–visible range during the plasma
polymerization of acetone/CO2 [164].
the relative concentration of species such as H, CH, CO, OH, and CN.
Examples of such analyses can also be found in the literature for ni-
trogen [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 technique consists in calculating the
line intensity ratios Ix/It of two plasma species. The first one, I x, is
related to the particle of unknown concentration, the other one, I t, is
emitted by the tracer (of known concentration). The excitation
threshold energy and the excitation cross section of these two emitters
must be nearly the same. In this way, the ratio of intensities Ix/It is
directly proportional to the concentration of the emitting species n x.
This line intensity ratio is equal to:

IX nXneKX
t ¼ t e t ; ð6Þ
I n n K

In Eq. (6), ne is the electron density and Kx, Kt are related to:
i) The production rate of the considered excited states
(considering electron impact excitation, the later depend on the
EEDF and the electron-specie collision excitation cross
section),
ii) The radiative lifetime of the considered excited species, and
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
iii) The global response of the optical setup (e.g., the transmission
emission intensity, (b) an increase in the deposition rate and (c) a decrease in the of the lenses, optical fibers, the sensitivity of the detector, etc.
sulfur content in the PPF. Inset: Typical emission spectrum of the ICP discharge taken de- pend on the wavelength).
in the inductive mode for a power of 100 W [132].

If the tracer t is properly chosen, the line ratio is proportional to


To the contrary, it is difficult to obtain quantitative data through nx. Since Kx, Kt, ne cancel out and nt is known, nx can be deduced. In Ref.
OES measurements, e.g., determining the absolute density of species [168], actinometry was implemented to quantify the production of H 2,
identi- fied from the data displayed in Fig.10. The detected intensity of H and CH species in an Ar/styrene plasma. The fragmentation of the
the line Ix, appearing at a wavelength λ, and emitted by an excited ar- omatic precursor was found to increase with the RF power
specie x can be expressed according to Eq. (5): 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 the
Ix ¼ nxω Aij kðλÞ; ð5Þ properties of the sty- rene – based PPF are improved at a RF power of
100 W. OES and film characterization data pointed out that
where nx⁎ is the number density of the excited species x and k(λ) improvement at this specific value of the power is due to the
characterizes the optical setup response at the wavelength λ. Aij is the predominance of CH radicals in the plas- ma and an enhanced cross-
frequency of spontaneous emission of a photon at the wavelength λ linking density due to the presence of a highest percentage of carbon
following the radiative decay from the excited state j toward the content in the film (Fig. 12). These re- searchers concluded their
lower energy state i. study by suggesting the possibility of using styrene-based PPF,
In cold low-pressure plasmas, the production of excited levels deposited at RF power of 100 W, as high perfor- mance protective
typi- cally involves electron collisions with the species. The rate coatings for metal surfaces. Using a similar methodol- ogy, Palumbo et
constant of this collision depends on the collision cross section σ(E), al. has highlighted an inverse correlation between the relative
which is a function of the kinetic energy of the electrons, and of the proportion of CO in the discharge and the retention degree of the
EEDF. The later is therefore a key property that should be precursor in acrylic acid PPF [167].
determined in order to extract quantitative information from the OES
measurements. Al- though one can find collision cross section data
in the literature, e.g., for elastic collisions with polyatomic molecules
relevant to plas- ma processing [202] and inelastic cross section for
electron with hy- drogen [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 ener- gy level, which is characterized by a much longer
radiative lifetime as compared to radiative excited states, another loss
mechanism, in- volving diffusion outside the detection area (e.g.,
toward the cham- ber walls), must also be taken into account. This
situation would make a quantitative description of the plasma even
more complicat- ed. This is the reason why OES should not be
considered as the tool of choice to provide a quantitative insight on
the plasma chemistry dur- ing plasma polymerization.
Nevertheless, in some special experimental conditions, actinom-
etry can be adapted to plasma polymerization in order to estimate
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].
In another study, Bousquet et al. used the actinometry method for plasma
investigating the temporal evolution of the relative density of atomic
oxygen by time-resolved OES measurements in HMDSO/O 2 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. Briefly, this method consists in ap-
plying 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 long-lived species. More details can be found in
Refs. [166,209]. This approach has allowed measuring the O-atom loss
coeffi- cient on surfaces as a function of the chemical composition of
the gas. Furthermore, it has been shown that deposition events during
the plas- ma “OFF” time occur due to the dissociation of HMDSO
molecules by long-lived oxygen atoms.
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
impact in the under- standing 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 technique for a
rapid screening of plasma chemistry.

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 emerged as the
most widely employed diagnostic method for investigating the plasma
polymerization process. Indeed, in comparison to the OES method, a
deeper knowledge of the chemical composition of the discharge
includ- ing ions, stable molecules and radicals can be obtained. The
first reports of the analysis of organic plasmas by MS date from the
1980s for poly- merizing/etching fluorine based-discharges
[197,210]. Nevertheless, the use of the mass 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. Short in Sheffield
[38,57,102,169,170].
In order to avoid: (i) failure of some electronic part of the
apparatus
and (ii) collisions of analyzed ions during their transport inside the
equipment, a pressure less than 5.10 −6 Torr is 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 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 detec-
tion of neutral (residual gas analysis, RGA mode) and ionic (glow dis-
charge 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 re-
quires an ionization source. The most common approach consists in
heating a filament to generate an electron beam normal to the 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 ion- ization energy of the molecules (typically 10 eV for
organic species). It should be noted that electron attachment
processes can also be employed leading to the 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 electronegative species (e.g.,
fluorine-based) and has therefore been significantly less employed.
We will therefore focus only on the electron impact ionization mode
for the detection of neutrals.
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
quad- rupole analyzer, the most commonly used mass analyzer for
analysis [175]. Its working principle relies on the combination of
DC and RF potentials applied to four conducting rods to define
stable trajec- tories in oscillating electrical fields which allow
separating ions accord- ing to their m/z ratio. It should be noted
that the quadrupole analyzer is characterized by a transmission
function which strongly decreases with mass. For most of mass
spectrometers, the manufacturer found empiri-
cally that this function 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 instrumenta- tion 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 cannot be obtained since the transit time of neutral
species across the plasma-instrument boundary layer and the
ionization chamber is typically longer than the pulse dura- tion
when the process is operated in pulsed mode [41].

4.3.1. Neutral species analysis


With regard to the RGA detection mode, a non-negligible
drawback is related to the ionization process taking place in the
ionization source of the equipment. Indeed, the ions are formed in
an excited state and the release of excess of energy may lead to the
fragmentation of the molec- ular 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 does not necessarily imply that the
corresponding species are formed in the plasma. In con- ventional
MS, a kinetic energy of 70 eV is usually employed in order to
maximize the signal intensity since for organic molecules, the
maxi- mum of the electron beam ionization cross section is located
around 70 eV [175,213]. Nevertheless, for plasma analysis, an
electron kinet- ic energy of 20 eV is preferable in order to limit
excessive precursor fragmentation in the ionization source of the
mass spectrometer [41,47,48,54,55,66,68,131]. This value
corresponds to the best com- promise between reduction of
fragmentation in the ionization source and the threshold
ionization energy of various organic-based molecules and
fragments [41].
Based on previous considerations, when operating in the RGA
mode,
the main challenge is then to differentiate the molecules/fragments
pro- duced through dissociation/rearrangement reactions in the
plasma from those formed in the ionization source of the
spectrometer. Al- though this aspect is often neglected, a very
simple relation (Eq. (7)) can be applied to subtract the
fragmentation undergone by the precur- sor in the apparatus itself
following [41,54,68,131]:
IMonomer ðPlasma ONÞ
I m I Plasma ON −I Plasma OFF 7
ð Þ¼ ð Þ ð Þ: ; ð
IMonomerðPlasma OFFÞ Þc m m

Fig. 13. Schematic description of the principle of mass spectrometry in RGA mode.
The 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 spectrum (in red) (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this
article.).
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 the precursor have been observed in some studies, no “dimer” or
“trimer” based on the 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 neutral dimer has been observed [215].
A closer look to the mass spectrum in Fig. 14 also reveals the
produc- tion of stable hydrocarbon based molecules (e.g. C 2H2, C2H4),
as fre- quently encountered whatever the precursor employed
[38,47,54,55, 104,128]. As proposed by Thiry et al. and supported by
theoretical calcu- lations, these hydrocarbon molecules result from
rearrangement reac- tions of the 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 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 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 cannot be trapped within
the PPF, these species do not take part in film growth, hence reducing
the amount of reactive molecules contain- ing 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 amu. This leads to numerous isobaric interferences
(i.e., several species contributing to the same ion signal). For 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 13C-labeled and
unlabeled acrylic acid in order to facilitate the attribution of the peaks
in the mass spectra [38]. In some particular cases, we can take
advantage of the isotopic abundance of the heteroelement to
discriminate molecules presenting similar m/z ratios, as shown in the
case of propanethiol dis- charges [66]. Finally, another strategy
consists in measuring the intensi- ty of the signal at a given m/z as a
Fig. 14. Neutral mass spectra of acrylic acid: (a) plasma OFF, (b) plasma ON (100 μs on- function of the kinetic energy of the colliding electrons [198]. This
time, 1000 μs off-time, 50 W), (c) corrected for both fragmentation of the acrylic acid method enables one to deduce the ioniza- tion energy of the specie(s)
precursor in the ionization source according to Eq. (7) and mass transmission [41]. contributing to a given signal and then to determine their chemical
composition. Ultimately, the combination of other diagnosis methods
such as gas phase Fourier transform infrared spectroscopy and OES
can also assist the interpretation of the mass spectra data [53,200].
where Ic (m) is the calculated peak intensity for m/z = m, Im Although the intensity of a signal in an RGA mass spectrum is related
(Plasma ON) and Im (Plasma OFF) represent the experimental peak to the density of the corresponding species, calculation of their
intensity for mass m when the plasma is switched ON and OFF, absolute concentration is not straightforward. Among other things,
respectively. IMonomer is the intensity corresponding to the precursor this requires the knowledge of the ionization cross-section which, for
signal. many stable molecules, are available in the literature [175]. In this
An illustrative example of the application of Eq. (7) is shown in case, the absolute density of the probed molecules can be obtained by
Fig. 14 for the plasma polymerization of acrylic acid. Together with means of a calibra- tion gas with a known density and electron impact
allylamine, this PPF is probably the most studied in the literature. ionization cross- section [175]. With respect to radical species,
Fig. 14 a-b present the mass spectra of acrylic acid (in the absence determination of their absolute density is much more demanding as
of plasma) and when the plasma is switched ON in pulsed mode they can be lost through collisions with the walls of the reactor
(50 W, ton = 100 μs, toff = 1000 μs), respectively. The peaks can be [173,175]. Most of the time, the corresponding cross-sections are not
assigned as follows [38,163]: m/z = 72 the acrylic acid precursor; available and several assumptions have therefore to be made [175].
m/z = 55 to CH2CHCO+; m/z = 44 to C H O .+
; C H .+
and CO .+
; Furthermore, in some cases, specific experimental configurations
3 82 4 2
m/z = 28 to such as multistep differential pumping
C H.+ and CO.+; m/z = 27 to C H+; m/z = 26 to C H.+; m/z = 18 to are needed for an accurate measurement [173,174]. Based on these
con-
2 4 2 3 2 2
H2O.+; m/z = 2 to H 2
.+
; and m/z = 1 to H+. The mass spectrum of siderations, in the context of the plasma polymerization process, the
the plasma corrected according to Eq. (7) is depicted in Fig. 14c. After great diversity of reactions (including fragmentations, rearrangements,
the correction, peaks previously identified at m/z = 55 and 27 almost etc.) highly complicates measurements of the absolute densities of the
disap- pear, indicating that these signals originate from dissociations numerous species constituting the plasma.
reactions in the ionization source. This example illustrates the Nevertheless, some simple methods can be employed for obtaining
importance of the correction method for obtaining the right picture of quantitative or semi-quantitative information. For instance, the abso-
the neutral plasma chemistry. lute precursor concentration can be easily measured using a
calibration
curve relying on the intensity of the monomer signal as a function of
present at the growing film surface. I rel. (H2S) has then been calculated
the pressure when the plasma is switched OFF [41]. Using this
following Eq. (9) and compared to the sulfur to the carbon ratio (S/C)
calibration curve, the intensity of the precursor measured in the mass
measured by XPS in the corresponding PPF (Fig. 15):
spectrome- ter, whatever the plasma parameters, can be directly
related to the par- tial pressure of the monomer. For example, a IC ðH2SÞ
I ðH SÞ ¼ ; ð9Þ
decrease in the partial
pressure of the acrylic acid precursor as a function of the energy rel: 2 X
m=z¼29 X
m=z¼43
deliv- ered in the plasma has been reported [37,38,41]. This expected ICðm=z ¼ 15Þ þ ICðm=zÞ þ IC ðm=zÞ
behavior is attributed to an increase in ne with power resulting in an m=z¼26 m=z¼39

increase in precursor fragmentation [142]. It should be noted that


the amount of
unfragmented chemical precursors remaining in the discharge is where Ic (m/z) corresponds to the corrected intensity calculated
gener- ally related to the concentration of the chemical group hosted follow- ing Eq. (7) for the signal at m/z.
by the monomer [47,48]. The obtained linear correlation between Irel. (H2S) and the S/C
To monitor the relative proportion of radicals and stable ratio as a function of power validates the trapping hypothesis,
molecules produced through fragmentation/rearrangement reactions revealing the attractiveness of the mass spectrometry technique for a
in the plas- ma, a depletion function can be employed according to Eq. deeper under- standing of the PPF growth at a molecular level.
(8) [172, 218]: We understand from these examples that the interpretation of
IOFF− ION ION mass spectrometry data is often challenging. As already mentioned,
D ð%Þ ¼ × 100 ¼ 1− × 100; ð8Þ one strat- egy to make it easier consists in exploiting DFT calculations.
As an exam-
IOFF IOFF ple, the study of the plasma polymerization of ethyl lactate, finding
applications in the design of biodegradable coatings, is presented [54].
where ION is the intensity of a peak at a given plasma condition and IOFF Briefly, the objective of this work was to correlate the ester content in
is the intensity of the same peak in the absence of a plasma. the films, a key parameter for controlling the biodegradability of
Based on Eq. (8), when a peak intensity increases with power, the the material, with the plasma chemistry. A particular feature of the
depletion of that peak is negative. Negative values indicate that the mass spectra of ethyl lactate plasma in RGA mode is the absence of a
cor- responding species are formed by the action of the plasma. In sig- nal corresponding to the precursor (m/z = 118), thus making the
other words, the higher the negative value of the depletion function, inter- pretation of the data highly complicated. This is explained by
the higher is the production of the corresponding species in the the “brittleness” of the ethyl lactate molecule which suffers of
plasma. Using this methodology, Hazrati et al. have studied the extended fragmentation reactions in the ionization source of the
formation of key fragments for the plasma polymerization of ethanol spectrometer. Based on mass spectrometry data, a thorough
and demon- strated the production of radicals such as .CH3 and .OH to mechanistic study using DFT was undertaken. A detailed description
the detriment of the precursor molecule upon increasing the power of the complex calculated fragmentation pathway occurring in the
[172]. It is impor- tant to stress here that Eq. (8) has to be employed ionization source of the spec- trometer as well as in the plasma (Fig.
with care depending on the plasma polymer investigated. Indeed, 16a–b) is beyond the scope of this review and interested readers can
species formed from the consult Ref. [54]. We will concen-
dissociation
affect of the precursor ion in the source of the spectrometer trate on the species at m/z = 75 (i.e., C H O+) which deserve peculiar
3 7 2
the measured intensities and therefore the conclusions drawn from 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.
the depletion function.
Another strategy has been recently implemented for directly
corre- lating plasma and film chemistries in view of the elucidation of
the growth mechanism of propanethiol PPF [68]. Briefly, under
certain experimental conditions, it has been shown that propanethiol
PPF con- tains 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
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 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
attention (see highlighted rectangle in Fig. 16). Using the DFT
calcula- tions, it can be learned that this species can only be
formed through an ionic mechanism involving the 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. 16 b), the
intermediate of the corresponding reaction is unstable and
rearranges spontaneously to a more stable form at m/ z = 74.
Comparing both fragmentation patterns (i.e., ionic and radical), this
implies that the intensity of the peak at m/z = 75 is proportional
to the density of undamaged precursor in the plasma. Therefore,
this peak can be correlated with the evolution of the ester content
in the PPF de- pending on the synthesis conditions. This example
highlights the signif- icance 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.
Another example revealing the strong interest in combining
DFT cal- culations 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 PPF [48]. It has been
shown that above a critical value of the mean power (around 30
W) 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
incorporation of primary amines in the films. However, below this
value, the content of amine 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 reactions associated with various initial
fragmentation schemes of allylamine and cyclopropylamine, as
calculated by DFT. Note that:
(i) all calculations have been performed in the unrestricted scheme
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].

and, (ii) activation and bond dissociation energies are almost 4.3.2. Ion analysis
equivalent for bond dissociation processes [221]. In the case of As already mentioned, the ions also play a major role in the PPF
allylamine, the more thermodynamically favorable reaction involves growth mechanism. The ionic gas phase chemistry has therefore also
the rupture of the C\\N bond to form allyl and NH2 radicals. The been extensively investigated by MS. Fig. 18 represents a typical
lower bond dissociation energy is rationalized by the stabilization of exam- ple of the chemical composition of positive ions in an acrylic
the allyl radical by resonance ef- acid discharge. Similar to the RGA analysis mode, numerous peaks are
fects. In contrast, the easiest reaction in the case of cyclopropylamine iden- tified. It is important to stress that in this case, the ions
is the opening of the ring, thus retaining the amine chemical during their transport in the mass spectrometer obviously do not
functional- ity on the precursor. This different behavior is fully suffer of fragmen- tation before their detection. Therefore, all peaks
consistent with the increased amount of amino group detected in the appearing in the spec- trum correspond to ions formed in the plasma.
film upon plasma po- lymerization of cyclopropylamine. Accounting The base peak at m/z = 73 is ascribed to the protonated monomer
for the entropic effects would systematically decrease by about 0.5 eV, ([M + H]+) while the most
the calculated bond dis- prominent fragments are observed at m/z = 39 (C H+), 55
sociation energies while keeping unchanged the previous conclusions.
3 3
The full consistency between the experimental and theoretical results (CH2CHCO+) and 57 (CH3CH2CO+). A relevant observation is the
thus paves the way toward a theoretical screening and even the detec- tion of oligomeric ions of the form (2 M + H)+ at m/z =
design of precursors aimed at optimization of the degree of PPF 145 and (3 M + H)+ at m/z = 217, where M corresponds to the
functionalization. molecular weight of the starting material as also observed in other
works [101,
Fig. 17. DFT-calculated bond dissociation energies of the initial fragmentations of allylamine (top) and cyclopropylamine (bottom). Adapted from [48].

128,215,222]. Oligomer ions corresponding to [4 M + H]+ have also


reason why neutral/radical addition reactions in the gas phase is
been identified in Ref. [38]. The dimer and trimer ions can lose H2O
ruled out lies in the kinetics of the reactions as proposed by Benedikt
giv- ing rise to additional signals separated by a value of 18. Such
oligomer ions formed through gas phase neutral/ion reactions have [226]. Indeed, ion-neutral reactions are typically 10 times faster than
also been identified in propanoic acid [215,222], allylalcohol [214], re- actions between neutral particles due to the attractive potential
propanol [214], allylamine [104,170], HMDSO [102,223,224], methyl created by the induced dipole moment in the neutral particle.
isobutyrate, methyl methacrylate, n-butyl methacrylate [169], ethanol The intensities of the peaks associated with the oligomeric cations
[172,225], γ-terpinene [171] and thiophene [218] plasmas. Most of were found to evolve with the power dissipated in the discharge.
the time, the formation of these oligomeric ions involves the addition Most of the time, a decrease in the relative proportion of these ions in
of a hydrogen ion on the neutral precursor followed by the successive the plasma upon increasing the power has been reported [37,169,
additions of un- charged monomers. However, the exact mechanism 171]. This behavior can be ascribed to a decrease in the concentration
remains unclear in some aspects. For example, gas phase of the precursor, hence reducing the probability of addition reactions
oligomerization reactions do not require the presence of a double with ions. Since a correlation between the relative proportion of ions
bond in the organic precursor marking a clear difference with and the degree of retention of the precursor in the films has been
classical ionic polymerization [169,215]. The highlighted for several PPF 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
af- terglow at sufficient long OFF time for enabling their extraction
from the bulk plasma. In continuous wave (CW) plasma
polymerization, their contribution in the condensing material is
excluded since they are con- fined 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
Fig. 18. Positive-ion mass spectrum of acrylic acid plasma sustained at a power of 3 W. plasma “OFF” times. Their role in the overall de- posited mass has been
The region under the horizontal bar has been expanded by a factor of 5 in intensity.
estimated to be quite low (b 1%) in comparison to positive ions and
Adapted from [215].
neutrals [228].
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].

ions bombarding the substrate is also of considerable interest for a


mechanistic description of plasma polymer formation. Indeed, by
means of a specific RF biasing technique enabling to control the ion
en- ergy independently of other 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 spectrometry technique allows us to
measure the energy dis- tribution function of the ions (IEDF) arriving
at the orifice of the instru- ment 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 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 estimation of the bombarding ion
energy, the orifice of the mass spec- trometer has to be localized at the
position of the substrate. Fig. 20 rep- resents a typical example of an
IEDF recorded in a γ-terpinene plasma using a spectrometer with a
grounded orifice. The IEDF has a shape which is typical of those
Fig. 19. Time-resolved ion mass spectra of acrylic acid discharges obtained at 50 W
recorded in plasmas characterized by a collisionless sheath. The
RF power (ton = 100 μs and toff = 10,000 μs). Adapted from [41]. “peak” in the 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)
Although the pulsed mode has been extensively employed in [171,225].
the plasma polymerization domain, only a few studies were dedicated Haddow et al. have performed a systematic study of the evolution
to time-resolved mass spectrometry measurements. As of the ion energy as a function of the dissipated power in CW plasma
aforementioned, such kinds of measurements are limited to ionic poly- merization of acrylic acid using an external RF inductive coil as
species [41]. The results obtained have allowed a better plasma source [38]. The results are summarized in Fig. 21 depicting an
understanding of the chemistry of pulsed organic discharges. The increase in the ion energy from 5 eV to 30 eV when the power evolves
main conclusions drawn are that some ions can “survive” during the from 1 to 15 W. At 3 W, the number of atoms in the molecular ion has
plasma “OFF” time during at least 1000 μs after the extinction of the been es- timated to be, on average, 10 whereas the ion energy is 13
plasma [41,163]. This aspect is illus- trated in Fig. 19 where GDMS eV, thus corresponding to an energy per atom b 1.3 eV [38,101]. It is
measurements were performed in acrylic acid discharges for selected reason- able to assume that at this value, fragmentation of the
times during the plasma “OFF” time. The dominant positive ions are striking ion (and surface) will be minimal enabling a better
identified as follows: m/z = 55 to CH2CHCO+, m/z = 73 to [M + H] preservation of the
+
(M being the acrylic acid precursor), m/z = 127 to [M + H-H2O]+,
m/z = 145 to [2 M + H]+, m/z = 199 to [3 M + H-H2O]+ and m/z =
217 to [3 M + H]+. After 500 μs, only the trimeric species (at m/z =
217) are still detected in the plasma and re- main 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.
In addition to the significance of determining the neutral/ionic
chemical composition of the plasma, investigating the energy of the Fig. 21. Evolution of the ion energy of acrylic acid cations relative to a self-biased, or
floating, surface as function of the power. Adapted from [38].
function of interest. At higher power, the peak ion energy is 30 eV
measurement of the ion flux using this probe, the readers are invited
while the ions contain b 10 atoms on average, leading to an energy
to consult the paper of Braithwate et al. [176]. Today, ion flux probes
per atom N 3 eV. Since bond energies in organic molecules typically
are relatively inexpensive and easy to use [58].
range from 3 to 5 eV, the fragmentation of the incoming ion and/or
An example of the evolution of the ion flux toward the substrate as
the growing film surface might be anticipated at higher powers,
a function of the power is displayed in Fig. 22 for the plasma
therefore contributing to the decrease in the degree of retention of polymeri- zation of allylamine. The ion flux was found to
the precursor in the coating [38,104].
increase from 6.6 × 1016 m−2 s−1 ions at 1 W to 1.4 × 10 18 ions m−2 s−1
To summarize, the popularity gained in the field of plasma
at 14 W. Sim- ilar trends and ion flux values were reported for acrylic
polymer- ization by the MS technique over several decades lies in its
acid [58,222],
ability to de- termine relevant properties including the chemical
propionic acid [58,222], hydrocarbons (n-hexane and 1,7-octadiene),
composition of the plasma and the energy of bombarding ions. This is diethylene glycol, diethylene glycol divinyl ether [58], γ-terpinene
probably the reason why this method has emerged as one of the most [171] and ethanol [225] plasmas. The observed trend in Fig. 22 is
powerful diagnostic methods for deepening the fundamental ascribed to the increase in the electron density and, in turn, in the ion
understanding of the mecha- nistic formation of a given PPF. The density in the plasma volume upon increasing the power [142]. The
main disadvantage of the MS tech- nique is related to the cost of the calculation of the average ion mass based on the ionic mass spectra
equipment and the often difficult interpretation of the data. In the combined with ion flux measurements allow us to deduce the ion
latter case, it is clear that theoretical support such as DFT calculations mass flux toward the growing film. To evaluate the contribution of
can help. ions in the deposited mass, the latter is compared with the mass
depo- sition rate (expressed as the total mass deposited per unit
4.4. Ion probes surface and time) obtained from quartz crystal microbalance. The
data are summa- rized in Table 2 for the present example. At low
Using the MS technique, although all ions constituting the plasma power (i.e., 1 W), ap- proximately 63% of ions could, in principle,
can be identified, the absolute ion flow toward the growing film contribute to film formation. Interestingly, at a power of 5 W, the ion
cannot be deduced since MS measurements do not give access to the flux is sufficient to ac- count for all deposited mass. This provides a
absolute density of ions. This aspect is, however, of crucial strong argument that ions significantly contribute to film formation. It
importance to deter- mine the mass delivery by ions to the film is important here to stress that it does not mean that the film grows
surface and also in combina- tion with other techniques to measure through a pure ionic mecha- nism and that neutral/radical surface
the contribution of ions in the total energy flux reaching the growing reactions are ruled out. Estimation of the mass delivery to the film by
film. In plasma processing, a classical means to measure the ion flux to ions by a comparison between the mass deposition rate and ion flux
a surface is based on ordinary electrostatic probes inserted into the inherently takes into account that all ionic species exhibit a sticking
plasma volume. From measure- ments of the ion density and electron coefficient equal to unity, or at least does not evolve with the process
temperature in the plasma bulk, the ion flux toward the interface is parameters. However, the real situation is much more complicated as
deduced using the Bohm criterion. Nevertheless, in the context of the sticking coefficient of ions is less than unity and is affected by
plasma polymerization, this method is not well suited for reliable and their chemical nature, energy and the nature of the surface. For
routine analysis. Indeed, as explained in section 4.1, the coverage of example, it can be learned from Table 2 that the ion flux exceeds the
the probe with an insulating PPF highly complicates the mass deposition rate for a power of 5 W, indicating a loss of ionic
measurements [8,101,104,176]. mass incorporated in the film. A possible loss mechanism could be
In 1996, Braithwaite et al. have developed a novel probe to recombination events occurring at the growing film surface followed
measure by the formation of a stable molecule unable to be chemisorbed
the absolute ion flux toward a surface in organic discharges which pre- (e.g., CO.+ + e− → CO (g) or CH CH CO+ + e− → CH CH
sents significant advantages: tolerant to insulating deposits, non- surface 3 2 surface 3 2

perturbing, easy to implement and applicable whatever the plasma + CO (g)) [101]. It should be noted that a more pronounced ab-
(surface)

source [176]. Briefly, the basic principle consists in applying a pulsed lation phenomenon at higher power might also be anticipated. The
RF voltage to the electrode. Owing to the different mobility of pre- vious considerations also prevail for the sticking coefficient of
electrons and ions, the probe naturally acquires a negative self-bias neutral species for which it has been reported that their surface
potential when an RF signal is applied. The RF signal is then chopped reactivity is di- rectly related to the density of dangling bonds at the
for approx- imately 5–10 ms and the change in the voltage is interface [99,107]. Actually, both neutrals and ions are intertwined in
measured with time as the ions impact the negatively biased electrode the overall mecha- nism since an increase in the ion flux causes the
[8,102]. The ion flux is deduced from the variation of the bias voltage formation of more rad- ical sites at the surface, thus resulting in a
as a function of time. For more details about the physical more efficient grafting of reactive neutral species.
phenomena involved in the The emergence of ion probes has also highlighted the influence of
the chemical nature of the precursor on the process. Michelmore et al.
investigated several saturated/unsaturated monomers, 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 pro- nounced ionic deposition mechanism for
the saturated precursor,

Table 2
Comparison of total mass deposition rates and positive-ion mass flux in allylamine
plasmas for different powers [104]. See text for details.

Fig. 22. Evolution of the ion flux in allylamine plasmas as a function of RF power [104].
Power (W) Total material deposition rate Positive–ion mass flux
(μg m−2 s−1) (μg m−2 s−1)

1 18.7 11.8
3 61.7 36.2
5 86 99.4
15 127.1 226.6
whereas the grafting of intact precursor at open bond sites present at the densities of species in the plasma. From an instrument point of
the interface is predominant for the unsaturated one. view, the sensitivity problem that is encountered in low-pressure dis-
The development of the ion probe has therefore allowed shedding charges is often tackled by implementing White-cell multiple pass
light on the contribution of the ions to the PPF formation. Even
though the exact role played by radicals and ions in the plasma
polymer forma- tion is still debated, it is however well-established
that the actual pic- ture of the PPF growth mechanism should
consider both species.

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 dis-
charges and, specifically, the plasma polymerization process
[41,48,68, 172]. Nevertheless, it is accepted that it presents two major
drawbacks, namely the fact that: (i) the quantification of species is
complex such that most of 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
mode makes interpretation of the data even more complicated [99].
These limitations are in fact related to the “ex situ” feature of this
plasma diagnostic technique. Therefore, in situ approaches have been
devel- oped in order to overcome these drawbacks while taking care
to avoid perturbing the plasma during measurements.
Optical techniques fulfill these requirements, especially the ab-
sorption spectroscopy (AS) method because they allow determining
population densities in both ground and metastable states as well as
information related to the gas temperature when considering the line
profile of atomic bands or the ro-vibrational structure of molecular
gases [231]. The fact that absolute densities can be deduced from AS
mea- surements without instrument calibration is one of the main
advantage of this technique in comparison to OES, which was discussed
in Section 4.2. Among the different wavelength regions that have been
scanned, the in-
frared (λ = 14 μm–25 μm corresponding to 700–4000 cm−1) is
particu-
larly well adapted to plasma polymerization process because this
region of the electromagnetic spectrum contains the vibrational
signatures of plasma-generated molecular species.
Historically, dispersive instruments were first used for IR
absorption measurements, but they are not well suited for detecting
weakly ab- sorbing molecular species in low pressure plasmas.
Therefore, today, plasma diagnostics using infrared absorption
spectroscopy is often per- formed through Fourier transform infrared
spectroscopy (FTIR) or using tuneable diode laser absorption
spectroscopy (TDLAS). Both methods have their advantages and
drawbacks. Laser techniques typically offer higher sensitivity and
resolution than FTIR, but cannot (even with tun- able lasers) probe
the broad bandwidth range required to track many IR absorption
peaks simultaneously. An excellent review on the TDLAS
spectroscopy can be found in Ref. [232].
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], using fluorocarbon gases to
take advantage of the strong absorbing nature of the fluorocarbon
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 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 molecu- lar and low- temperature
plasmas characteristic of the plasma polymer- ization 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 re- action
kinetics. This is only possible if we are able to quantitatively probe
optical arrangements. Fig. 23 illustrates, for example, typical experi- the more complex plasma polymerization systems that are today
mental setups used for FTIR (Fig. 23a) and TDLAS (Fig. 23b) studied to synthesize functionalized organic surfaces. As an example,
measurements. Wells et al. have studied the plasma polymerization of ethylene glycol
Numerous PECVD processes involving hydrocarbons precursors in a pulsed inductively-coupled plasma reactor by FTIR and OES in
to deposit C-containing films such as Diamond Like Carbon layers or order to
plas- ma polymers have been studied by IR absorption spectroscopy
in order to monitor transient and stable species that are strongly
involved 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].
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 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
con- sistent with the higher growth rate and fluorine content
obtained when using the C5F8 precursors [236]. Fig. 24 shows the
power depen- dence 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 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 thermal
stability of the films in comparison with pure C 5F8 films [236]. The
PECVD deposition of C-containing coatings with other
heteroelements than fluorine 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 relatively high pressure (i.e., 1 Torr) by using
OES and FTIR (without the use of a White cell) [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 degree of
fragmentation of the HDMSO precursor as a function of the injected
power, which was 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 FTIR absorption spectroscopy [178]. In this case, the
plasma was gen- erated at low pressure (3 mTorr) using a Withe cell
with up to 44 m of optical length (see Fig. 23a). These data were
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 rate in the films. Although a
signifi-
cant 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 pre- senting numerous bands related to the different
chemical species/radi- cals in the plasma makes interpretation
difficult.
This later claim becomes even more relevant when considering
Fig. 23. 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].

evaluate the dynamics of monomer fragmentation and the effective


spectrometry section [53]. This approach has been developed in the
chemical feedback from the boundary walls [59]. A pressure of
context of the synthesis of ethyl lactate plasma polymers. In order to
60 mTorr was used and the optical length was 64 cm. The presence in
validate our approach, we have first compared the infrared spectrum
the plasma of CH, CO, OH and H (by OES) and of acetylene, ethylene,
simulated for ethyl lactate with a corresponding experimental mea-
methane, water, formaldehyde, CO and CO2 (by FTIR) was
surement obtained from ethyl lactate vapor (plasma OFF), see Fig. 26.
demonstrat- ed. These stable volatile molecules are claimed to be the
A comparison between the two spectra shows a very nice agreement,
results of radi- cal recombination and polymerization processes
with a root-mean-square deviation of ~ 5 cm−1, hence validating our
occurring at plasma/ walls interfaces. It is also shown that evolution
theoretical approach.
of the film chemistry as a function of the applied power is correlated
When initiating the plasma, the fragmentation of precursor mole-
with the fragmentation pathway of ethylene glycol in pulsed plasmas
cules generates a mixture of different species leading to a much more
with a higher retention of the monomer functionality at low mean
complex spectrum. Mass spectrometry measurements in RGA mode
power.
have allowed identifying nine dominant fragments: CH 3, CH4, H2O, CO,
The complexity of plasma polymerization processes clearly makes
CO2, C2H4, C2H2, CH3CHOH, and C3H7O2. Fig. 27b exhibits the sum of
infrared data interpretation more and more difficult. In order to
the DFT-calculated infrared spectra of these nine fragments, with
contrib- ute to a better interpretation of these data, our group has
their relative intensities weighted on the basis of the peak intensities
recently used a DFT strategy similar to the one mentioned
from
earlier in the mass
Fig. 24. Power dependence of (a) CF, (b) CF2 and (c) CF3 radical densities measured by TDLAS for different fluorocarbon-based precursors [235].

the mass spectrum. The weighting factor (WFi) for a given species i is
synthetic IR spectrum. This can be corrected by estimating the
expressed as:
percent- age of undamaged ethyl lactate molecules in the plasma (X EL)
by com- paring the intensity of the peaks related to the ester
IMS
function in the
ethyl lactate vapor when the plasma is switched OFF and when a
WFi ¼ X
i
i
:
ð Þ power of 60 W is applied. This leads to a XEL value equal to 0.712. As a
i MS
I

with IiMS the intensity in the mass spectrum of the species i.


This yields a simulated spectrum in marked discrepancy with the
ex- perimental one (Fig. 27a) due to 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

Fig 25. Simplified reaction pathway for (a) low and (b) high power conditions Fig. 26. (a) Experimental IR spectrum of ethyl lactate in gas phase without discharge at
established from FTIR diagnostic measurements for the plasma polymerization of 10 mTorr and 5 sccm and (b) IR spectrum of ethyl lactate calculated by DFT [53].
HMDSO. Adapted from [178].
specifically, to evaluate the reaction efficiency of the ester-containing
fragments (σester) in the growing film [99]. The latter was obtained by
comparing the plasma chemistry to the film composition. Fig. 28
shows the evolution of σester as a function of the applied power. The
measured trend highlights the higher incorporation efficiency of
ester- based fragments in the growing film at high powers. This is
likely relat- ed to a higher density of surface active site generated by
ion and photon irradiation upon increasing the power.
Gas-phase FTIR spectroscopy is a powerful tool allowing for
quanti- tative determination of important species (radical, stable
molecules) in the plasma phase. In combination with other diagnostic
tools, as well as with theoretical calculations and surface analysis, it is
even possible to evaluate surface reaction coefficients. One of the
important drawbacks is the complexity of the instrumentation which
requires the design of dedicated reactors.

5.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”.
Never- theless, 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 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 chemistry employed in low-pressure
Fig. 27. (a) Experimental IR spectrum of an ethyl lactate plasma sustained by a power of plasma polymerization processes can only be obtained by a
60 W. (b) Synthetic IR spectrum calculated on the basis of Eq. (10) and (c) corrected complementary ap- proach combining several state-of-the-art plasma
according to the degree of fragmentation of the precursor [53]. diagnostic tools. The most employed tools are mass spectrometry (in
both ion and neutral modes), optical techniques such as OES and gas
result, the peaks associated to the fragments are further corrected by phase FTIR, and to a lesser extent electrostatic and ionic probes. Each
a factor (1 − XEL) = 0.288. The resulting corrected spectrum, shown in of these techniques provides a piece of a puzzle that has to be handled
Fig. 27c, appears to be in much better agreement with the recorded with care. Indeed, in addition to the already challenging experimental
ex- measurements, the interpretation of the gathered data is, at least, as
perimental spectrum, allowing for the assignment of most of the difficult. Therefore, in order to access the physical values that will
peaks even though the presence of fragments containing ester groups ultimately be used to de- termine a convincing growth mechanism
in the plasma has been neglected. This original protocol clearly (e.g., flux of radicals, ions, sticking coefficients, etc.), the use of
provides a new diagnostic tool to probe the plasma chemistry, theoretical calculations such as those based on DFT can provide
especially the pres- ence of specific chemical functionalities. valuable support.
In the framework of this research, we were also interested in It is clear that the research community has understood the strong
the evolution of the ester-containing species density as a function of need for detailed knowledge of plasma chemistry as revealed by the
the power which, in fine, partly determine the degradation behavior nu- merous publications on this topic. Nevertheless, effort is still
of the deposited films. These data were used to better understand the required to go further in the direction of a quantitative description of
growth mechanism of the ethyl lactate plasma polymer and, more this complex 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 coat- ing application. The wise choice and even the
design of precursors that can be driven by diagnostic tools and
theoretical calculation as demon- strated in some work reported in
this review are a good examples of the future knowledge-driven
development of the plasma polymerization process.
The future will also certainly see the confirmation of the rise of at-
mospheric 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
this technology. Therefore, it is clear that diagnostic approaches,
Fig. 28. Evolution of the surface reaction coefficient of ester-bearing fragments (σester) as
a function of power during the synthesis of ethyl lactate plasma polymers, as estimated specific to the atmospheric pressure plasma polymerization
from FTIR plasma diagnostic data. Adapted from [99]. processes, should also be on the “to-do” list of the scientists active in
the field.
Altogether, we are convinced that large avenues are open for the [25] A. Felten, C. Bittencourt, J.-J. Pireaux, G. Van Lier, J.-C. Charlier, Radio-frequency
use of plasma polymerization processes, but the price to pay will be plasma functionalization of carbon nanotubes surface O 2, NH3, and CF4 treatments,
more and more precise understanding of fundamental mechanisms J. Appl. Phys. 98 (2005) 074308.
[26] Q. Chen, L. Dai, M. Gao, S. Huang, A. Mau, Plasma activation of carbon nanotubes for
which will only be obtained by a thorough evaluation of the plasma
chemical modification, J. Phys. Chem. B 105 (2001) 618–622.
chemistry using a combination of diagnostic tools. [27] B. Akhavan, K. Jarvis, P. Majewski, Plasma polymerization of sulfur-rich and water-
stable coatings on silica particles, Surf. Coat. Technol. 264 (2015) 72–79.
[28] B. Akhavan, K. Jarvis, P. Majewski, Plasma polymer-functionalized silica particles
Acknowledgments for heavy metals removal, ACS Appl. Mater. Interfaces 7 (2015) 4265–4274.
[29] T.D. Michl, B.R. Coad, A. Hü sler, K. Vasilev, H.J. Griesser, Laboratory scale systems
for the plasma treatment and coating of particles, Plasma Process. Polym. 12
D. Thiry thanks the “Région Pays de La Loire (France)” through (2015) 305–313.
the “Post-doctorats internationaux” program and the FRIA grant [30] K.L. Jarvis, P. Majewski, Influence of particle mass and flow rate on plasma poly-
from the “Fédé ration Wallonie-Bruxelles” for financial support. S. merized allylamine coated quartz particles for humic acid removal, Plasma
Process. Polym. 12 (2015) 42–50.
Konstantinidis and J. Cornil are research associate and director of re- [31] A. Choukourov, I. Melnichuk, A. Shelemin, P. Solař, J. Hanus, D. Slavinska, H.
search of the Fonds National de la Recherche Scientifique (FNRS) Biederman, Plasma polymerization on mesoporous surfaces: n-hexane on
titanium
Belgium, respectively. D. Thiry, S. Konstantinidis and R. Snyders
nano-particles, J. Phys. Chem. C (2015), http://dx.doi.org/10.1021/acs.jpcc.5b08604.
acknowl- edge the support of the Belgian Government (Belspo) [32] V. Barranco, J. Carpentier, G. Grundmeier, Correlation of morphology and
through the “Pô le d'attraction interuniversitaire” (PAI, P07/14, barrier properties of thin microwave plasma polymer films on metal substrate,
“Plasma-Surface Interac- tion”, ψ) and of Wallonia Region (DG06) Electrochim. Acta 49 (2004) 1999–2013.
[33] G. Grundmeier, P. Thiemann, J. Carpentier, V. Barranco, Tailored thin plasma polymers
through the program of excel- lence “OPTI2MAT” and the for the corrosion protection of metals, Surf. Coat. Technol. 174–175 (2003) 996–
CONVERGENCE project “EVERWALL”. 1001.
[34] M. Deilmann, S. Theiß, P. Awakowicz, Pulsed microwave plasma polymerization of
silicon oxide films: application of efficient permeation barriers on polyethylene
References terephthalate, Surf. Coat. Technol. 202 (2008) 1911–1917.
[35] J. Schneider, D. Kiesler, M. Leins, A. Schulz, M. Walker, U. Schumacher, U. Stroth,
Development of plasma polymerised SiOx barriers on polymer films for food pack-
[1] K. Bewilogua, G. Brä uer, A. Dietz, J. Gä bler, G. Goch, B. Karpuschewski, B. Szyszka, aging applications, Plasma Process. Polym. 4 (2007) S155–S159.
Surface technology for automotive engineering, CIRP Ann. Manuf. Technol. 58 [36] M.R. Alexander, T.M. Duc, A study of the interaction of acrylic acid/1,7-
(2009) 608–627. octadiene plasma deposits with water and other solvents, Polymer 40 (1999)
[2] C.D. Dimitrakopoulos, D.J. Mascaro, Organic thin-film transistors: a review of 5479–5488.
re- cent advances, IBM J. Res. Dev. 45 (2001) 11–27. [37] S. Fraser, R.D. Short, D. Barton, J.W. Bradley, A multi-technique investigation of the
[3] M. Ohring, Materials Science of Thin Films, Elsevier Science, USA, 2001. pulsed plasma and plasma polymers of acrylic acid: millisecond pulse regime, J.
[4] H. Biederman, Y. Osada, Plasma Technology, Elsevier Science Publishers, Phys. Chem. B 106 (2002) 5596–5603.
Amsterdam, 1992. [38] D.B. Haddow, R.M. France, R.D. Short, J.W. Bradley, D. Barton, A mass
[5] A. Hemberg, S. Konstantinidis, F. Renaux, J.P. Dauchot, R. Snyders, Ion flux–film spectrometric and ion energy study of the continuous wave plasma
structure relationship during magnetron sputtering of WO 3, Eur. Phys. J. Appl. polymerization of acrylic acid, Langmuir 16 (2000) 5654–5660.
Phys. 56 (2011) 24016. [39] D. Hegemann, E. Kö rner, K. Albrecht, U. Schü tz, S. Guimond, Growth mechanism
[6] K. Sarakinos, J. Alami, S. Konstantinidis, High power pulsed magnetron of oxygen-containing functional plasma polymers, Plasma Process. Polym. 7
sputtering: a review on scientific and engineering state of the art, Surf. Coat. (2010) 889–898.
Technol. 204 (2010) 1661–1684. [40] D. Hegemann, U. Schü tz, E. Kö rner, Macroscopic approach to plasma
[7] P.J. Kelly, R.D. Arnell, Magnetron sputtering: a review of recent developments polymeriza- tion using the concept of energy density, Plasma Process. Polym. 8
and applications, Vacuum 56 (2000) 159–172. (2011) 689–694.
[8] A. Michelmore, D.A. Steele, J.D. Whittle, J.W. Bradley, R.D. Short, Nanoscale [41] S.A. Voronin, M. Zelzer, C. Fotea, M.R. Alexander, J.W. Bradley, Pulsed and
deposi- tion of chemically functionalised films via plasma polymerisation, RSC continu- ous wave acrylic acid radio frequency plasma deposits: plasma and
Adv. 3 (2013) 13540–13557. surface chem- istry, J. Phys. Chem. B 111 (2007) 3419–3429.
[9] N.J. Saikia, C. Ewels, J.-F. Colomer, B. Aleman, M. Amati, L. Gregoratti, A. Hemberg, [42] C.L. Rinsch, X. Chen, V. Panchalingam, R.C. Eberhart, J.-H. Wang, R.B. Timmons,
D. Thiry, R. Snyders, C. Bittencourt, Plasma fluorination of vertically aligned Pulsed radio frequency plasma polymerization of allyl alcohol: controlled
carbon nanotubes, J. Phys. Chem. C 117 (2013) 14635–14641. deposi- tion of surface hydroxyl groups, Langmuir 12 (1996) 2995–3002.
[10] P. Chabert, N. Braithwaite, Physics of Radio-Frequency Plasmas, Cambridge [43] L. Watkins, A. Bismarck, A.F. Lee, D. Wilson, K. Wilson, An XPS study of pulsed plas-
Univer- sity Press, New York, 2011. ma polymerised allyl alcohol film growth on polyurethane, Appl. Surf. Sci. 252
[11] A. Grill, Cold Plasma Materials Fabrication: from Fundamentals to (2006) 8203–8211.
Applications, Wiley, New York, 1994. [44] A. Choukourov, H. Biederman, I. Kholodkov, D. Slavinska, M. Trchova, A. Hollander,
[12] K. Vasilev, J. Cook, H.J. Griesser, Antibacterial surfaces for biomedical devices, Ex- Properties of amine-containing coatings prepared by plasma polymerization, J.
pert Rev. Med. Devices 6 (2009) 553–567. Appl. Polym. Sci. 92 (2004) 979–990.
[13] B.T. Houseman, E.S. Gawalt, M. Mrksich, Maleimide-functionalized self- [45] A. Choukourov, H. Biederman, D. Slavinska, L. Hanley, A. Grinevich, H. Boldyryeva,
assembled monolayers for the preparation of peptide and carbohydrate A. Mackova, Mechanistic studies of plasma polymerization of allylamine, J.
biochips†, Langmuir 19 (2002) 1522–1531. Phys. Chem. B 109 (2005) 23086–23095.
[14] P. Jonkheijm, D. Weinrich, H. Schrö der, C.M. Niemeyer, H. Waldmann, Chemical [46] D. Debarnot, T. Mé rian, F. Poncin-Epaillard, Film chemistry control and growth
strategies for generating protein biochips, Angew. Chem. Int. Ed. 47 (2008) ki- netics of pulsed plasma-polymerized aniline, Plasma Chem. Plasma Process.
9618–9647. 31 (2010) 217–231.
[15] A. Misra, P. Dwivedi, Immobilization of oligonucleotides on glass surface using [47] L. Denis, D. Cossement, T. Godfroid, F. Renaux, C. Bittencourt, R. Snyders, M. Hecq,
an efficient heterobifunctional reagent through maleimide–thiol combination Synthesis of allylamine plasma polymer films: correlation between plasma diag-
chem- istry, Anal. Biochem. 369 (2007) 248–255. nostic and film characteristics, Plasma Process. Polym. 6 (2009) 199–208.
[16] E.A. Smith, M.J. Wanat, Y. Cheng, S.V.P. Barreira, A.G. Frutos, R.M. Corn, [48] L. Denis, P. Marsal, Y. Olivier, T. Godfroid, R. Lazzaroni, M. Hecq, J. Cornil, R.
Formation, spectroscopic characterization, and application of sulfhydryl- Snyders, Deposition of functional organic thin films by pulsed plasma polymeriza-
terminated alkanethiol monolayers for the chemical attachment of DNA onto tion: a joint theoretical and experimental study, Plasma Process. Polym. 7
gold surfaces, Langmuir 17 (2001) 2502–2507. (2010) 172–181.
[17] H. Biederman, Introducion, in: H. Biederman (Ed.), Plasma Polymer Films, [49] K. Vasilev, L. Britcher, A. Casanal, H.J. Griesser, Solvent-induced porosity in
Imperial College Press, London 2004, pp. 15–16. ultrathin amine plasma polymer coatings, J. Phys. Chem. B 112 (2008) 10915–
[18] N. Bondt, J.R. Deimann, A.P.V. Trostwijk, J. Fourcroy Ann. Chem. 21 (1796) 58. 10921.
[19] J. Goodman, The formation of thin polymer films in the gas discharge, J. Polym. Sci. [50] C. Daunton, L.E. Smith, J.D. Whittle, R.D. Short, D.A. Steele, A. Michelmore,
44 (1960) 551–552. Plasma parameter aspects in the fabrication of stable amine functionalized
[20] J. Friedrich, The Plasma Chemistry of Polymer Surfaces, Wiley, Weinhein, 2012. plasma poly- mer films, Plasma Process. Polym. 12 (2015) 817–826.
[21] J. Friedrich, Mechanisms of plasma polymerization—reviewed from a chemical [51] L. Denis, D. Thiry, D. Cossement, P. Gerbaux, F. Brusciotti, I. Van De Keere, V.
point of view, Plasma Process. Polym. 8 (2011) 783–802. Goossens, H. Terryn, M. Hecq, R. Snyders, Towards the understanding of
[22] N. Inagaki, Plasma Surface Modification and Plasma Polymerization, Technomic plasma polymer film behaviour in ethanol: a multi-technique investigation,
Publishing, Lancaster, 1996. Prog. Org. Coat. 70 (2011) 134–141.
[23] L. Jorge, S. Coulombe, P.-L. Girard-Lauriault, Nanofluids containing MWCNTs [52] S. Ligot, F. Renaux, L. Denis, D. Cossement, N. Nuns, P. Dubois, R. Snyders,
coated with nitrogen-rich plasma polymer films for CO 2 absorption in aqueous Experi- mental study of the plasma polymerization of ethyl lactate, Plasma
medium, Plasma Process. Polym. 12 (2015) 1311–1321. Process. Polym. 10 (2013) 999–1009.
[24] D. Shi, J. Lian, P. He, L. Wang, W.J. van Ooij, M. Schulz, Y. Liu, D.B. Mast, Plasma [53] S. Ligot, M. Guillaume, P. Raynaud, D. Thiry, V. Lemaur, T. Silva, N. Britun, J. Cornil,
de- position of ultrathin polymer films on carbon nanotubes, Appl. Phys. Lett. P. Dubois, R. Snyders, Experimental and theoretical study of the plasma
81 (2002) 5216–5218. chemistry of ethyl lactate plasma polymerization discharges, Plasma Process.
Polym. 12 (2015) 405–415.
[54] S. Ligot, M. Guillaume, P. Gerbaux, D. Thiry, F. Renaux, J. Cornil, P. Dubois, R.
Snyders, Combining mass spectrometry diagnostic and density functional theory
calculations for a better understanding of the plasma polymerization of Ethyl lac-
tate, J. Phys. Chem. B 118 (2014) 4201–4211. [85] K. Vasilev, N. Poulter, P. Martinek, H.J. Griesser, Controlled release of levofloxacin
[55] L. Denis, F. Renaux, D. Cossement, C. Bittencourt, N. Tuccitto, A. Licciardello, M. sandwiched between two plasma polymerized layers on a solid carrier, ACS
Hecq, R. Snyders, Physico-chemical characterization of methyl isobutyrate-based Appl. Mater. Interfaces 3 (2011) 4831–4836.
plasma polymer films, Plasma Process. Polym. 8 (2011) 127–137. [86] K. Vasilev, Z. Poh, K. Kant, J. Chan, A. Michelmore, D. Losic, Tailoring the surface
[56] T.R. Gengenbach, H.J. Griesser, Deposition conditions influence the functionalities of titania nanotube arrays, Biomaterials 31 (2010) 532–540.
postdeposition oxidation of methyl methacrylate plasma polymer films, J. [87] S. Simovic, D. Losic, K. Vasilev, Controlled drug release from porous materials
Polym. Sci. A Polym. Chem. 36 (1998) 985–1000. by plasma polymer deposition, Chem. Commun. 46 (2010) 1317–1319.
[57] A. Michelmore, P. Gross-Kosche, S.A. Al-Bataineh, J.D. Whittle, R.D. Short, On the ef- [88] O. Kyliá n, A. Choukourov, H. Biederman, Nanostructured plasma polymers,
fect of monomer chemistry on growth mechanisms of nonfouling PEG-like plasma Thin Solid Films 548 (2013) 1–17.
polymers, Langmuir 29 (2013) 2595–2601. [89] P.L. Girard-Lauriault, W.E. Unger, P.M. Dietrich, A. Hollä nder, Innovative and
[58] A. Michelmore, D.A. Steele, D.E. Robinson, J.D. Whittle, R.D. Short, The link between established strategies for the surface analysis of nitrogen and oxygen-rich plasma
mechanisms of deposition and the physico-chemical properties of plasma polymer polymer films by XPS: an introductory guide, Plasma Process. Polym. 12 (2015)
films, Soft Matter 9 (2013) 6167–6175. 953–967.
[59] G.P. Wells, I.C. Estrada-Raygoza, P.L.S. Thamban, C.T. Nelson, C.-W. Chung, L.J. [90] B. Nisol, F. Reniers, Challenges in the characterization of plasma polymers using
Overzet, M.J. Goeckner, Understanding the synthesis of ethylene glycol pulsed XPS, J. Electron Spectrosc. Relat. Phenom. 200 (2015) 311–331.
plasma discharges, Plasma Process. Polym. 10 (2013) 119–135. [91] A. Bogaerts, E. Neyts, R. Gijbels, J. van der Mullen, Gas discharge plasmas and their
[60] R. D'Agostino, F. Cramarossa, F. Fracassi, E. Desimoni, L. Sabbatini, P.G. Zambonin, applications, Spectrochim. Acta B At. Spectrosc. 57 (2002) 609–658.
[92] M.A. Lieberman, A.J. Lichtenberg, Principle of Plasma Discharge and Materials Pro-
G. Caporiccio, Polymer film formation in C 2F6-H2 discharges, Thin Solid Films 143
cessing, Wiley, New York, 2005.
(1986) 163–175.
[93] A. Fridman, Plasma Chemistry, Cambridge University Press, New York, 2008.
[61] P. Favia, Plasma deposition of fluoropolymer films in different glow discharges re-
gimes, in: H. Bierderman (Ed.), Plasma Polymer Films, Imperial College Press, [94] D. Merche, N. Vandencasteele, F. Reniers, Atmospheric plasmas for thin film
London 2004, pp. 46–47. depo- sition: a critical review, Thin Solid Films 520 (2012) 4219–4236.
[62] F. Henry, F. Renaux, S. Coppé e, R. Lazzaroni, N. Vandencasteele, F. Reniers, R. [95] S. Guimond, U. Schü tz, B. Hanselmann, E. Kö rner, D. Hegemann, Influence of gas
phase and surface reactions on plasma polymerization, Surf. Coat. Technol. 205
Snyders, Synthesis of superhydrophobic PTFE-like thin films by self-
(Supplement 2) (2011) S447–S450.
nanostructuration in a hybrid plasma process, Surf. Sci. 606 (2012) 1825–1829.
[96] H. Yasuda, Luminous Chemical Vapor Deposition and Interface, Marcel Dekker,
[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, Lang- New York, 2005.
muir 26 (2009) 3388–3393. [97] H. Yasuda, Plasma Polymerization, Elsevier Science, 1985.
[64] W.C.E. Schofield, J. McGettrick, T.J. Bradley, J.P.S. Badyal, S. Przyborski, Rewritable [98] A. Milella, F. Palumbo, P. Favia, G. Cicala, R. d'Agostino, Continuous and modulated
DNA microarrays, J. Am. Chem. Soc. 128 (2006) 2280–2285. deposition of fluorocarbon films from c-C4F8 plasmas, Plasma Process. Polym. 1
[65] L.G. Harris, W.C.E. Schofield, K.J. Doores, B.G. Davis, J.P.S. Badyal, Rewritable (2004) 164–170.
glycochips, J. Am. Chem. Soc. 131 (2009) 7755–7761. [99] S. Ligot, D. Thiry, P.A. Cormier, P. Raynaud, P. Dubois, R. Snyders, In situ IR
[66] D. Thiry, N. Britun, S. Konstantinidis, J.-P. Dauchot, M. Guillaume, J. Cornil, R. spectros- copy as a tool to better understand the growth mechanisms of plasma
Snyders, Experimental and theoretical study of the effect of the inductive-to- polymers thin films, Plasma Process. Polym. 12 (2015) 1200–1207.
ca- pacitive transition in propanethiol plasma polymer chemistry, J. Phys. [100] A. Michelmore, J.D. Whittle, R.D. Short, The importance of ions in low pressure
Chem. C 117 (2013) 9843–9851. PECVD plasmas, Front. Phys. 3 (2015) 3.
[67] D. Thiry, R. Francq, D. Cossement, D. Guerin, D. Vuillaume, R. Snyders, [101] S. Candan, A.J. Beck, L. O'Toole, R.D. Short, A. Goodyear, N. St, J. Braithwaite, The
Establish- ment of a derivatization method To quantify thiol function in sulfur- role of ions in the continuous-wave plasma polymerisation of acrylic acid, Phys.
containing plasma polymer films, Langmuir 29 (2013) 13183–13189. Chem. Chem. Phys. 1 (1999) 3117–3121.
[68] D. Thiry, R. Francq, D. Cossement, M. Guillaume, J. Cornil, R. Snyders, A detailed de- [102] A. Michelmore, P.M. Bryant, D.A. Steele, K. Vasilev, J.W. Bradley, R.D. Short, Role
scription of the chemistry of thiol supporting plasma polymer films, Plasma of positive ions in determining the deposition rate and film chemistry of
Pro- cess. Polym. 11 (2014) 606–615. continuous wave hexamethyl disiloxane plasmas, Langmuir 27 (2011) 11943–
[69] L.M.H. Groenewoud, G.H.M. Engbers, J. Feijen, Plasma polymerization of thiophene 11950.
derivatives, Langmuir 19 (2003) 1368–1374. [103] A. Michelmore, C. Charles, R.W. Boswell, R.D. Short, J.D. Whittle, Defining
[70] L.M.H. Groenewoud, G.H.M. Engbers, J.G.A. Terlingen, H. Wormeester, J. Feijen, plasma polymerization: new insight into what we should be measuring, ACS
Pulsed plasma polymerization of thiophene, Langmuir 16 (2000) 6278–6286. Appl. Mater. Interfaces 5 (2013) 5387–5391.
[71] S. Liu, M.M.L.M. Vareiro, S. Fraser, A.T.A. Jenkins, Control of attachment of [104] A.J. Beck, S. Candan, R.D. Short, A. Goodyear, N.S.J. Braithwaite, The role of ions
bovine serum albumin to pulse plasma-polymerized maleic anhydride by in the plasma polymerization of allylamine, J. Phys. Chem. B 105 (2001) 5730–
variation of pulse conditions, Langmuir 21 (2005) 8572–8575. 5736.
[72] K.S. Siow, L. Britcher, S. Kumar, H.J. Griesser, Plasma methods for the generation of [105] P.N. Brookes, S. Fraser, R.D. Short, L. Hanley, E. Fuoco, A. Roberts, S. Hutton, The ef-
fect of ion energy on the chemistry of air-aged polymer films grown from the
+
hyperthermal polyatomic ion Si2OMe 5 , J. Electron Spectrosc. Relat. Phenom.
121
chemically reactive surfaces for biomolecule immobilization and cell colonization coatings—as a biodegradable solid carrier for tunable drug delivery applications,
—a review, Plasma Process. Polym. 3 (2006) 392–418. Polymer 54 (2013) 4820–4829.
[73] M.N. Macgregor-Ramiasa, A.A. Cavallaro, K. Vasilev, Properties and reactivity
of polyoxazoline plasma polymer films, J. Mater. Chem. 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 Surf. B: Biointerfaces 126
(2015) 70–79.
[75] B.R. Coad, M. Jasieniak, S.S. Griesser, H.J. Griesser, Controlled covalent surface
im- mobilisation of proteins and peptides using plasma methods, Surf. Coat.
Technol. 233 (2013) 169–177.
[76] A. Manakhov, P. Sklá dal, D. Nečas, J. Čechal, J. Polčá k, M. Eliá š, L. Zajíčková ,
Cyclopropylamine plasma polymers deposited onto quartz crystal
microbalance
for biosensing application, Phys. Status Solidi (a) 211 (2014) 2801–2808.
[77] K. Bazaka, M. Jacob, W. Chrzanowski, K. Ostrikov, Anti-bacterial surfaces:
natural agents, 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 nanocoatings for bio-applications, Surg. Innov. 3 (2015) 63–83.
[79] P. Qi, Y. Yang, K.-Q. Xiong, Q. Tu, Z. Yang, J. Wang, J. Chen, N. Huang, J. Wang,
Mul- tifunctional plasma polymerized film: towards better anti-corrosion
property, en- hanced cellular growth ability, attenuated inflammatory and
histological responses, ACS Biomater. Sci. Eng. 1 (2015) 513–524.
[80] J.W. Haycock, 3D cell culture: a review of current approaches and techniques,
in:
J.W. Haycock (Ed.), 3D Cell Culture, Springer 2011, pp. 1–15.
[81] J.F. Friedrich, R. Mix, G. Kü hn, Functional groups bearing plasma homo and
copol- ymer layers as adhesion promoters in metal–polymer composites, Surf.
Coat. Technol. 174-175 (2003) 811–815.
[82] R. Fö rch, 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.
[84] S. Bhatt, J. Pulpytel, M. Mirshahi, F. Arefi-Khonsari, Plasma co-polymerized nano
(2001) 281–297.
[106] A. Choukourov, J. Kousal, D. Slavıńská , H. Biederman, E.R. Fuoco, S.
Tepavcevic, J. Saucedo, L. Hanley, Growth of primary and secondary amine
films from polyatom- ic ion deposition, Vacuum 75 (2004) 195–205.
[107] A. Von Keudell, T. Schwarz-Selinger, M. Meier, W. Jacob, Direct identification
of the synergism between methyl radicals and atomic hydrogen during
growth of amor- phous hydrogenated carbon films, Appl. Phys. Lett. 76
(2000) 676–678.
[108] P. Traskelin, O. Saresoja, K. Nordlund, Molecular dynamics simulations of
C2, C2H, C2H2, C2H3, C2H4, C2H5, and C2H6 bombardment of diamond (111)
surfaces, J. Nucl. Mater. 375 (2008) 270–274.
[109] A. Von Keudell, Surface processes during thin-film growth, Plasma
Sources Sci. Technol. 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
hydro- carbon precursor on the mechanical properties of a-C:H coatings
synthesized by a hybrid PECVD/PVD method, Plasma Process. Polym.
(2015), http://dx.doi.org/10. 1002/ppap.201500050.
[111] J. Robertson, Diamond-like amorphous carbon, Mater. Sci. Eng. R. Rep. 37
(2002) 129–281.
[112] J.D. Whittle, D.A. Steele, R.D. Short, Reconciling the physical and chemical
environ- ments of plasma: a commentary on “Mechanisms of plasma
polymerisation— reviewed from a chemical point of view”, Plasma Process.
Polym. 9 (2012) 840–843.
[113] R. d'Agostino, F. Palumbo, Comment on “ion-assisted processes of
polymerization in low-pressure plasmas”, Plasma Process. Polym. 9 (2012)
844–849.
[114] T.R. Gengenbach, Z.R. Vasic, R.C. Chatelier, H.J. Griesser, A multi-technique
study of the spontaneous oxidation of N-hexane plasma polymers, J.
Polym. Sci. A Polym. Chem. 32 (1994) 1399–1414.
[115] R.C. Chatelier, X. Xie, T.R. Gengenbach, H.J. Griesser, Quantitative analysis of
poly- mer 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 al- cohol, acrylic acid, allylamine, and octa-1,7-diene plasma
polymers as studied by X-ray photoelectron spectroscopy, Chem. Mater.
12 (2000) 2664–2671.
[117] S. Ershov, F. Khelifa, P. Dubois, R. Snyders, Derivatization of free radicals in
an isopropanol plasma polymer film: the first step toward polymer grafting,
ACS Appl. Mater. Interfaces 5 (2013) 4216–4223.
[118] P.-L. Girard-Lauriault, P.M. Dietrich, T. Gross, T. Wirth, W.E.S. Unger, Chemical
char- acterization of the long-term ageing of nitrogen-rich plasma polymer films [147] D. Hegemann, H. Brunner, C. Oehr, Deposition rate and three-dimensional uni-
under various ambient conditions, Plasma Process. Polym. 10 (2013) 388–395. formity of RF plasma deposited SiOx films, Surf. Coat. Technol. 142 (2001)
[119] T.R. Gengenbach, H.J. Griesser, Aging of 1,3-diaminopropane plasma-deposited 849–855.
polymer films: mechanisms and reaction pathways, J. Polym. Sci. A Polym. [148] D. Hegemann, E. Koerner, S. Guimond, Plasma polymerization of acrylic acid
Chem. 37 (1999) 2191–2206. revisited, Plasma Process. Polym. 6 (2009) 246–254.
[120] S. Ershov, F. Khelifa, M.-E. Druart, Y. Habibi, M.-G. Olivier, R. Snyders, P. Dubois, [149] D. Hegemann, Macroscopic control of plasma polymerization processes, Pure Appl.
Free radical-induced grafting from plasma polymers for the synthesis of thin Chem. 80 (2008) 1893–1900.
bar- rier coatings, RSC Adv. 5 (2015) 14256–14265. [150] D. Hegemann, M.M. Hossain, Influence of non-polymerizable gases added
[121] F. Khelifa, S. Ershov, M.-E. Druart, Y. Habibi, D. Chicot, M.-G. Olivier, R. Snyders, P. during plasma polymerization, Plasma Process. Polym. 2 (2005) 554–562.
Dubois, A multilayer coating with optimized properties for corrosion protection [151] D. Hegemann, E. Kö rner, S. Guimond, Reply to: “Testing the hypothesis:
of Al, J. Mater. Chem. A 3 (2015) 15977–15985. comments on plasma polymerization of acrylic acid revisited”, Plasma Process.
[122] D. Teare, W. Schofield, R. Garrod, J. Badyal, Rapid polymer brush growth by Polym. 7 (2010) 371–375.
TEMPO-mediated controlled free-radical polymerization from swollen plasma de- [152] R.D. Short, D.A. Steele, Testing the hypothesis: comments on Plasma
posited poly (maleic anhydride) initiator surfaces, Langmuir 21 (2005) polymerisa- tion of acrylic acid revisited, Plasma Process. Polym. 7 (2010)
10818–10824. 366–370.
[123] D. Teare, W. Schofield, V. Roucoules, J. Badyal, Substrate-independent growth [153] H. Biederman, O. Kyliá n, Some remarks to macroscopic kinetics of plasma
of micropatterned polymer brushes, Langmuir 19 (2003) 2398–2403. polymer- ization, Plasma Process. Polym. 8 (2011) 475–477.
[124] F. Bé nard, P. Dubois, M. Olivier, R. Snyders, L. Denis, F. Khelifa, D. Thiry, F. [154] A. von Keudell, J. Benedikt, A physicist's perspective on “Views on macroscopic ki-
netics of plasma polymerisation”, Plasma Process. Polym. 7 (2010) 376–379.
Renaux, Grafted Polymer Coatings, University of Mons and Materia Nova,
[155] R.B. Timmons, A.J. Griggs, Plasma Polymer Films, in: H. Bierderman (Ed.), Plasma
WO2011092212A1, 2011.
Polymer Films, Imperial College Press, London 2004, pp. 217–245.
[125] Y. Yin, K. Fisher, N.J. Nosworthy, D. Bax, R.J. Clarke, D.R. McKenzie, M.M. Bilek,
[156] R. Jafari, M. Tatoulian, W. Morscheidt, F. Arefi-Khonsari, Stable plasma
Com- parison on protein adsorption properties of diamond-like carbon and
nitrogen- containing plasma polymer surfaces, Thin Solid Films 520 (2012) polymerized acrylic acid coating deposited on polyethylene (PE) films in a low
frequency dis- charge (70 kHz), React. Funct. Polym. 66 (2006) 1757–1765.
3021–3025.
[157] D. Hegemann, E. Kö rner, N. Blanchard, M. Drabik, S. Guimond, Densification of
[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 functional plasma polymers by momentum transfer during film growth, Appl.
(2010) 708–714. Phys. Lett. 101 (2012) 211603.
[127] L.-Q. Chu, W. Knoll, R. Fö rch, Stabilization of plasma-polymerized allylamine [158] D. Hegemann, B. Hanselmann, S. Guimond, G. Fortunato, M.-N. Giraud, A.G.
Guex, Considering the degradation effects of amino-functional plasma polymer
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 plas- coatings for biomedical application, Surf. Coat. Technol. 255 (2014) 90–95.
[159] A. Choukourov, H. Biederman, D. Slavinska, M. Trchova, A. Hollander, The
ma deposition: acrylic acid revisited, J. Vac. Sci. Technol. A 16 (1998) 1702–1709.
influence of pulse parameters on film composition during pulsed plasma
[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 polymerization of diaminocyclohexane, Surf. Coat. Technol. 174–175 (2003)
effect of addition of 1,7-octadiene to plasma deposits of acrylic acid, Chem. 863–866.
Mater. 12 (2000) 866–868. [160] T. Mé rian, D. Debarnot, F. Poncin-Epaillard, Effect of fluorine substitution of
[130] S. Swaraj, U. Oran, A. Lippitz, J.F. Friedrich, W.E.S. Unger, Aging of plasma- aniline ring on pulsed plasma polymer growth and structure, Plasma Process.
deposited films prepared from organic monomers, Plasma Process. Polym. 4 Polym. 8 (2011) 763–772.
(2007) S784–S789. [161] G. Mishra, S.L. McArthur, Plasma polymerization of maleic anhydride: just what
[131] D. Thiry, F.J. Aparicio, N. Britun, R. Snyders, Concomitant effects of the substrate are the right deposition conditions? Langmuir 26 (2010) 9645–9658.
temperature and the plasma chemistry on the chemical properties of [162] M. Dhayal, J.W. Bradley, Time-resolved electric probe measurements in the pulsed-
propanethiol plasma polymer prepared by ICP discharges, Surf. Coat. Technol. plasma polymerisation of acrylic acid, Surf. Coat. Technol. 194 (2005) 167–174.
[163] S. Voronin, M. Alexander, J. Bradley, Time-resolved mass and energy spectral
241 (2014) 2–7.
inves- tigation of a pulsed polymerising plasma struck in acrylic acid, Surf. Coat.
[132] D. Thiry, N. Britun, S. Konstantinidis, J.-P. Dauchot, L. Denis, R. Snyders, Altering
Technol. 201 (2006) 768–775.
the sulfur content in the propanethiol plasma polymers using the capacitive-
[164] M. Drabik, C. Celma, J. Kousal, H. Biederman, D. Hegemann, Properties of a C:H:O
to- inductive mode transition in inductively coupled plasma discharge, Appl.
plasma polymer films deposited from acetone vapors, Thin Solid Films 573
Phys. Lett. 100 (2012) 071604.
(2014) 27–32.
[133] B. Finke, K. Schrö der, A. Ohl, Structure retention and water stability of
[165] A. Bousquet, A. Granier, G. Cartry, A. Goullet, Kinetics of O and H atoms in
microwave plasma polymerized films from allylamine and acrylic acid, Plasma
pulsed O2/HMDSO low pressure PECVD plasmas, J. Optoelectron. Adv. Mater.
Process. Polym. 6 (2009) S70–S74.
10 (2008) 1999–2002.
[134] D.E. Robinson, D.J. Buttle, J.D. Whittle, K.L. Parry, R.D. Short, D.A. Steele, The sub-
[166] A. Bousquet, G. Cartry, A. Granier, Investigation of O-atom kinetics in O2, CO2,
strate and composition dependence of plasma polymer stability, Plasma Process.
H2O and O2/HMDSO low pressure radiofrequency pulsed plasmas by time-
Polym. 7 (2010) 102–106.
resolved op- tical emission spectroscopy, Plasma Sources Sci. Technol. 16
[135] S. Swaraj, U. Oran, A. Lippitz, J.F. Friedrich, W.E.S. Unger, Surface analysis of
(2007) 597.
plasma- deposited polymer films, 6, Plasma Process. Polym. 2 (2005) 572–580.
[167] F. Palumbo, P. Favia, A. Rinaldi, M. Vulpio, R. d'Agostino, PE-CVD of organic
[136] Q. Chen, R. Fö rch, W. Knoll, Characterization of Pulsed plasma polymerization
allylamine as an adhesion layer for DNA adsorption/hybridization, Chem. Mater. thin films with controlled surface concentration of carboxylic groups, Plasmas
Polym. 4 (1999) 133–145.
16 (2004) 614–620.
[137] Z. Zhang, Q. Chen, W. Knoll, R. Foerch, R. Holcomb, D. Roitman, Plasma polymer [168] A. Choudhury, J. Chutia, S. Barve, H. Kakati, A. Pal, N. Mithal, R. Kishore, M. Pandey,
film structure and DNA probe immobilization, Macromolecules 36 (2003) D. Patil, Studies of physical and chemical properties of styrene-based plasma poly-
mer films deposited by radiofrequency Ar/styrene glow discharge, Prog. Org. Coat.
7689–7694.
70 (2011) 75–82.
[138] D. Szmigiel, C. Hibert, A. Bertsch, E. Pamuła, K. Domański, P. Grabiec, P. Prokaryn,
[169] L. O'Toole, R.D. Short, A.P. Ameen, F.R. Jones, Mass spectrometry of and
A. Ścisłowska-Czarnecka, B. Płytycz, Fluorine-based plasma treatment of
deposition- rate measurements from radiofrequency-induced plasmas of methyl
biocompati- ble silicone elastomer: the effect of temperature on etch rate and
surface proper- isobutyrate, methyl methacrylate and n-butyl methacrylate, J. Chem. Soc. Faraday
ties, Plasma Process. Polym. 5 (2008) 246–255. Trans. 91 (1995) 1363–1370.
[139] T.B. Casserly, K.K. Gleason, Effect of substrate temperature on the plasma poly- [170] A.J. Beck, S. Candan, R.M. France, F.R. Jones, R.D. Short, A mass spectral
merization of poly(methyl methacrylate), Chem. Vap. Depos. 12 (2006) 59– investigation of the RF plasmas of small organic compounds: an investigation of
66. the plasma- phase reactions in the plasma deposition from allyl amine,
[140] D. Cossement, F. Renaux, D. Thiry, S. Ligot, R. Francq, R. Snyders, Chemical and mi- Plasmas Polym. 3 (1998) 97–114.
crostructural characterizations of plasma polymer films by time-of-flight second- [171] J. Ahmad, K. Bazaka, J.D. Whittle, A. Michelmore, M.V. Jacob, Structural
ary ion mass spectrometry and principal component analysis, Appl. Surf. Sci. 355 characteri- zation of γ-terpinene thin films using mass spectroscopy and X-ray
(2015) 842–848. photoelectron spectroscopy, Plasma Process. Polym. 12 (2015) 1085–1094.
[141] J.D. Whittle, R.D. Short, D.A. Steele, J.W. Bradley, P.M. Bryant, F. Jan, H. [172] H.D. Hazrati, J.D. Whittle, K. Vasilev, A mechanistic study of the plasma
Biederman, polymeri- zation of ethanol, Plasma Process. Polym. 11 (2014) 149–157.
A.A. Serov, A. Choukurov, A.L. Hook, Variability in plasma polymerization [173] H. Singh, J.W. Coburn, D.B. Graves, Mass spectrometric detection of reactive
processes—an international round-robin study, Plasma Process. Polym. 10 (2013) neutral species: beam-to-background ratio, J. Vac. Sci. Technol. A 17 (1999)
767–778. 2447–2455.
[142] M. Dhayal, J.W. Bradley, Using heated probes in plasma polymerising [174] J. Benedikt, S. Agarwal, D. Eijkman, W. Vandamme, M. Creatore, M. Van de
discharges, Surf. Coat. Technol. 184 (2004) 116–122. Sanden, Threshold ionization mass spectrometry of reactive species in remote
Ar∕C2H2 expanding thermal plasma, J. Vac. Sci. Technol. A 23 (2005) 1400–
[143] M.A. Gilliam, Q. Yu, H. Yasuda, Plasma polymerization behavior of fluorocarbon
monomers in low-pressure AF and RF discharges, Plasma Process. Polym. 4 1412.
(2007) 165–172. [175] J. Benedikt, A. Hecimovic, D. Ellerweg, A. Von Keudell, Quadrupole mass spectrom-
[144] D. Hegemann, Macroscopic investigation of reaction rates yielding plasma etry of reactive plasmas, J. Phys. D. Appl. Phys. 45 (2012) 403001.
[176] N.S.J. Braithwaite, J. Booth, G. Cunge, A novel electrostatic probe method for ion
polymer deposition, J. Phys. D. Appl. Phys. 46 (2013) 205204.
flux measurements, Plasma Sources Sci. Technol. 5 (1996) 677.
[145] D. Hegemann, Plasma polymerization and its applications in textiles, Indian J.
Fibre Text. Res. 31 (2006) 99. [177] P. Raynaud, T. Amilis, Y. Segui, Infrared absorption analysis of organosilicon/oxygen
plasmas in a microwave multipolar plasma excited by distributed electron cyclo-
[146] D. Hegemann, M.M. Hossain, E. Kö rner, D.J. Balazs, Macroscopic description of
tron resonance, Appl. Surf. Sci. 138 (1999) 285–291.
plas- ma polymerization, Plasma Process. Polym. 4 (2007) 229–238.
[178] P. Raynaud, B. Despax, Y. Segui, H. Caquineau, FTIR plasma phase analysis of
hexamethyldisiloxane discharge in microwave multipolar plasma at different elec-
trical powers, Plasma Process. Polym. 2 (2005) 45–52.
[179] A. Bogaerts, M. Eckert, M. Mao, E. Neyts, Computer modelling of the plasma
chem- istry and plasma-based growth mechanisms for nanostructured materials, [208] S.F. Durrant, M.A.B. De Moraes, PECVD of amorphous hydrogenated
J. Phys. D. Appl. Phys. 44 (2011) 174030. oxygenated nitrogenated carbon films, J. Polym. Sci. B Polym. Phys. 36 (1998)
[180] A.C. Van Duin, S. Dasgupta, F. Lorant, W.A. Goddard, ReaxFF: a reactive force field 1881–1888.
for hydrocarbons, J. Phys. Chem. A 105 (2001) 9396–9409. [209] A. Bouchoule, P. Ranson, Study of volume and surface processes in low pressure
[181] R. Car, M. Parrinello, Unified approach for molecular dynamics and density- radio frequency plasma reactors by pulsed excitation methods. I. Hydrogen–
functional theory, Phys. Rev. Lett. 55 (1985) 2471. argon plasma, J. Vac. Sci. Technol. A 9 (1991) 317–326.
[210] G. Turban, B. Grolleau, P. Launay, P. Briaud, A mass spectrometric diagnostic of
[182] A. Cheesman, J.N. Harvey, M.N. Ashfold, Studies of carbon incorporation on the di-
C2F6 and CHF3 plasmas during etching of SiO 2 and Si, Revue de physique appliquée
amond {100} surface during chemical vapor deposition using density functional
20 (1985) 609–620.
theory, J. Phys. Chem. A 112 (2008) 11436–11448.
[211] F. Moix, K. McKay, J.L. Walsh, J.W. Bradley, Atmospheric-pressure plasma polymer-
[183] T.B. Casserly, K.K. Gleason, Chemical vapor deposition of organosilicon thin
ization of acrylic acid: gas-phase ion chemistry, Plasma Process. Polym. (2015),
films from methylmethoxysilanes, Plasma Process. Polym. 2 (2005) 679–687.
http://dx.doi.org/10.1002/ppap.201500031.
[184] K. Fukui, Formulation of the reaction coordinate, J. Phys. Chem. 74 (1970) 4161–
4163. [212] E. Stoffels, W. Stoffels, K. Tachibana, Electron attachment mass spectrometry as
[185] E.C. Yang, X.J. Zhao, P. Tian, J.K. Hao, Density functional theory and MP2 a diagnostics for electronegative gases and plasmas, Rev. Sci. Instrum. 69
calcula- tions of the transition states and reaction paths on coupling reaction of (1998) 116–122.
methane through plasma, Chin. J. Chem. 22 (2004) 430–433. [213] E. de Hoffmann, V. Stroobant, Mass Spectrometry: Principles and Applications,
[186] L.A. Curtiss, K. Raghavachari, P.C. Redfern, J.A. Pople, Assessment of Gaussian-2 Wiley, Chichester, 2013.
and density functional theories for the computation of enthalpies of formation, J. [214] S. Candan, Radio frequency-induced plasma polymerization of allyl alcohol and 1-
Chem. Phys. 106 (1997) 1063–1079. propanol, Turk. J. Chem. 26 (2002) 783–792.
[187] O. Auciello, D.L. Flamm, Plasma Diagnostics: Surface Analysis and Interactions, [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
Elsevier Science, San Diego, 2013.
[188] H. Biederman, P. Hlídek, J. Zemek, D. Slavínská , J. Jezek, P. Zakouril, J. Glosik, Depo- Trans. 91 (1995) 3907–3912.
sition and properties of hydrophilic films prepared by plasma polymerization of [216] C.-H. Tsai, W.-J. Lee, C.-Y. Chen, P.-J. Tsai, G.-C. Fang, M. Shih, Difference in
conver- sions between dimethyl sulfide and methanethiol in a cold plasma
Ar/n-hexane/H2O, Vacuum 46 (1995) 1413–1418.
[189] A. Brockhaus, G. Leu, V. Selenin, K. Tarnev, J. Engemann, Electron release in the environment, Plasma Chem. Plasma Process. 23 (2003) 141–157.
[217] M. Alexander, T. Duc, The chemistry of deposits formed from acrylic acid plasmas,
af- terglow of a pulsed inductively-coupled radiofrequency oxygen plasma,
J. Mater. Chem. 8 (1998) 937–943.
Plasma Sources Sci. Technol. 15 (2006) 171.
[218] B. Mitu, V. Satulu, G. Dinescu, Mass spectrometry diagnostic during RF plasma po-
[190] J. Greene, Optical spectroscopy for diagnostics and process control during glow
lymerization of thiophene vapors, Rom. J. Physiol. 56 (2011) 120–125.
dis- charge etching and sputter deposition, J. Vac. Sci. Technol. 15 (1978) 1718–
1729. [219] D. Thiry, F.J. Aparicio, P. Laha, H. Terryn, R. Snyders, Surface temperature: a key
[191] J. Greene, F. Sequeda-Osorio, B. Natarajan, Glow discharge optical spectroscopy for pa- rameter to control the propanethiol plasma polymer chemistry, J. Vac. Sci.
microvolume elemental analysis, J. Appl. Phys. 46 (1975) 2701–2709. Technol. A 32 (2014) 050602.
[220] S. Ershov, F. Khelifa, V. Lemaur, J. Cornil, D. Cossement, Y. Habibi, P. Dubois, R.
[192] A. Manakhov, L. Zajíčková, M. Eliá š, J. Čechal, J. Polčá k, J. Hnilica, Š. Bittnerová ,
D. Nečas, Optimization of cyclopropylamine plasma polymerization toward Snyders, Free radical generation and concentration in a plasma polymer: the effect
enhanced layer stability in contact with water, Plasma Process. Polym. 11 of aromaticity, ACS Appl. Mater. Interfaces 6 (2014) 12395–12405.
(2014) 532–544. [221] P. Marsal, M. Roche, P. Tordo, P. De Sainte Claire, Thermal stability of OH and
[193] M. Buddhadasa, P.-L. Girard-Lauriault, Plasma co-polymerisation of ethylene, 1, O- alkyl bonds in N-alkoxyamines. A density functional theory approach, J.
3- butadiene and ammonia mixtures: amine content and water stability, Thin Phys. Chem. A 103 (1999) 2899–2905.
Solid Films 591 (2015) 76–85. [222] D. Barton, A.G. Shard, R.D. Short, J.W. Bradley, The effect of positive ion energy
[194] A. Choukourov, H. Biederman, I. Kholodkov, D. Slavinska, M. Trchova, A. Hollander, on plasma polymerization: a comparison between acrylic and propionic acids, J.
Properties of amine-containing coatings prepared by plasma polymerization, J. Phys. Chem. B 109 (2005) 3207–3211.
Appl. Polym. Sci. 92 (2004) 979–990. [223] M.R. Alexander, F.R. Jones, R.D. Short, Mass spectral investigation of the radio-
[195] A. Choukourov, H. Biederman, D. Slavinska, M. Trchova, A. Hollander, The frequency plasma deposition of hexamethyldisiloxane, J. Phys. Chem. B 101
influence of pulse parameters on film composition during pulsed plasma (1997) 3614–3619.
polymerization of diaminocyclohexane, Surf. Coat. Technol. 174 (2003) 863– [224] M. Alexander, F. Jones, R. Short, Radio-frequency hexamethyldisiloxane plasma de-
866. position: a comparison of plasma-and deposit-chemistry, Plasmas Polym. 2 (1997)
[196] H. Yasuda, T. Yasuda, The competitive ablation and polymerization (CAP) principle 277–300.
and the plasma sensitivity of elements in plasma polymerization and treatment, [225] S. Saboohi, M. Jasieniak, B.R. Coad, H.J. Griesser, R.D. Short, A. Michelmore,
J. Polym. Sci. A Polym. Chem. 38 (2000) 943–953. Compar- ison of plasma polymerization under collisional and collision-less
[197] R. d'Agostino, F. Cramarossa, S. De Benedictis, Diagnostics and decomposition pressure re- gimes, J. Phys. Chem. B (2015),
mechanism in radio-frequency discharges of fluorocarbons utilized for plasma http://dx.doi.org/10.1021/acs.jpcb.5b07309.
etching or polymerization, Plasma Chem. Plasma Process. 2 (1982) 213–231. [226] J. Benedikt, Plasma-chemical reactions: low pressure acetylene plasmas, J. Phys. D.
[198] K. Aumaille, A. Granier, M. Schmidt, B. Grolleau, C. Vallee, G. Turban, Study of oxy- Appl. Phys. 43 (2010) 043001.
gen/tetraethoxysilane plasmas in a helicon reactor using optical emission [227] A. Howling, L. Sansonnens, J.L. Dorier, C. Hollenstein, Time-resolved
spectros- copy and mass spectrometry, Plasma Sources Sci. Technol. 9 (2000) measurements of highly polymerized negative ions in radio frequency silane
331. plasma deposition experiments, J. Appl. Phys. 75 (1994) 1340–1353.
[199] A. Granier, M. Vervloet, K. Aumaille, C. Vallée, Optical emission spectra of TEOS [228] I. Swindells, S.A. Voronin, P.M. Bryant, M.R. Alexander, J.W. Bradley, Temporal evo-
and HMDSO derived plasmas used for thin film deposition, Plasma Sources Sci. lution of an electron-free afterglow in the pulsed plasma polymerisation of acrylic
Technol. 12 (2003) 89. acid, J. Phys. Chem. B 112 (2008) 3938–3947.
[200] M. Goujon, T. Belmonte, G. Henrion, OES and FTIR diagnostics of HMDSO/O 2 gas [229] I. Swindells, S.A. Voronin, C. Fotea, M.R. Alexander, J.W. Bradley, Detection of
mixtures for SiOx deposition assisted by RF plasma, Surf. Coat. Technol. 188 neg- ative molecular ions in acrylic acid plasma: some implications for
(2004) 756–761. polymerization mechanisms, J. Phys. Chem. B 111 (2007) 8720–8722.
[201] N. Benissad, C. Boisse-Laporte, C. Vallée, A. Granier, A. Goullet, Silicon dioxide [230] D. Barton, R.D. Short, S. Fraser, J.W. Bradley, The effect of ion energy upon
depo- sition in a microwave plasma reactor, Surf. Coat. Technol. 116 (1999) plasma polymerization deposition rate for acrylic acid, Chem. Commun. 348-349
868–873. (2003).
[202] J.-S. Yoon, M.-Y. Song, H. Kato, M. Hoshino, H. Tanaka, M.J. Brunger, S. Buckman, [231] T. Cleland, D. Hess, In situ FTIR diagnostics of the radio-frequency plasma
H. Cho, Elastic cross sections for electron collisions with molecules relevant to decom- position of N2O, Plasma Chem. Plasma Process. 7 (1987) 379–394.
plasma processing, J. Phys. Chem. Ref. Data 39 (2010) 033106. [232] J. Rö pcke, G. Lombardi, A. Rousseau, P. Davies, Application of mid-infrared
[203] J.-S. Yoon, M.-Y. Song, J.-M. Han, S.H. Hwang, W.-S. Chang, B. Lee, Y. Itikawa, tuneable diode laser absorption spectroscopy to plasma diagnostics: a review,
Cross sections for electron collisions with hydrogen molecules, J. Phys. Chem. Plasma Sources Sci. Technol. 15 (2006) S148.
Ref. Data 37 (2008) 913–931. [233] J. Nishizawa, N. Hayasaka, In situ observation of plasmas for dry etching by IR
[204] Y. Itikawa, A. Ichimura, K. Onda, K. Sakimoto, K. Takayanagi, Y. Hatano, M. Hayashi, spec- troscopy and probe methods, Thin Solid Films 92 (1982) 189–198.
H. Nishimura, S. Tsurubuchi, Cross sections for collisions of electrons and photons [234] C. Mogab, A. Adams, D.L. Flamm, Plasma etching of Si and SiO 2—the effect of oxy-
with oxygen molecules, J. Phys. Chem. Ref. Data 18 (1989) 23–42. gen additions to CF4 plasmas, J. Appl. Phys. 49 (1978) 3796–3803.
[205] S.F. Durrant, R. Landers, G.G. Kleiman, S.G. Castro, M.A.B. de Moraes, Fluorine- [235] K. Takahashi, A. Itoh, T. Nakamura, K. Tachibana, Radical kinetics for polymer film
containing amorphous hydrogenated carbon films, Thin Solid Films 281 deposition in fluorocarbon (C 4 F8, C3 F6 and C5 F8) plasmas, Thin Solid Films 374
(1996) 294–297. (2000) 303–310.
[206] S.F. Durrant, M.A.B. de Moraes, Dynamic actinometric optical emission spectrosco- [236] T. Shirafuji, A. Tsuchino, T. Nakamura, K. Tachibana, Plasma copolymerization
py for the elucidation of plasma processes in the production of fluorinated of C6F6/C5F8 for application of low-dielectric-constant fluorinated amorphous
amor- phous hydrogenated carbon films from glow discharges, Thin Solid carbon films and its gas-phase diagnostics using in situ Fourier transform
Films 277 (1996) 115–120. infrared spec- troscopy, Jpn. J. Appl. Phys. 43 (2004) 2697.
[207] S.F. Durrant, M.A.B. de Moraes, Conventional and dynamic actinometry of glow
dis- charges fed mixtures of tetramethylsilane, sulfur hexafluoride, and helium,
J. Vac. Sci. Technol. A 16 (1998) 509–513.
View publication stats

You might also like