Freire 2003

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Amorphous Hydrogenated Carbon Films Deposited by PECVD: Nitrogen

Incorporation during Film Growth and by Plasma Surface Processing


F. L. Freire

Citation: AIP Conference Proceedings 669, 354 (2003); doi: 10.1063/1.1593938


View online: https://doi.org/10.1063/1.1593938
View Table of Contents: http://aip.scitation.org/toc/apc/669/1
Published by the American Institute of Physics

Articles you may be interested in


PECVD growth of carbon nanotubes: From experiment to simulation
Journal of Vacuum Science & Technology B, Nanotechnology and Microelectronics: Materials, Processing,
Measurement, and Phenomena 30, 030803 (2012); 10.1116/1.3702806

Effects of deposition temperature on the properties of hydrogenated tetrahedral amorphous carbon


Journal of Applied Physics 82, 4566 (1997); 10.1063/1.366193

Application of amorphous carbon based materials as antireflective coatings on crystalline silicon solar cells
Journal of Applied Physics 110, 043510 (2011); 10.1063/1.3622515
Amorphous Hydrogenated Carbon Films Deposited by
PECVD: Nitrogen Incorporation during Film Growth and
by Plasma Surface Processing
F. L. Freire Jr.

Departamento de Física, Pontifícia Universidade Católica do Rio de Janeiro


Caixa Postal 38071, 22453-970, Rio de Janeiro, RJ, Brazil

Abstract. Amorphous carbon films are being currently used in a wide variety of applications. The properties of the films
can be tuned by the deposition technique employed and by the growth conditions. One way to improve these properties is
through the incorporation into the amorphous skeleton of several elements, like H, N, F and Si. In this work, we review
the effects of nitrogen incorporation into hydrogenated carbon films (a-C:H) deposited by plasma enhanced chemical
vapor deposition (PECVD) and by post-deposition nitrogen plasma processing of a-C:H films. Modifications on film
microstructure, surface morphology, mechanical and tribological properties are discussed.

INTRODUCTION
Amorphous carbon films (a-C) are presently being used in a wide variety of applications as protective coatings:
automotive components, shaving blades, biomedical implants and computer hard disks [1]. The use of hydrogenated
carbon films (a-C:H) as the interconnect dielectric in ultra-large scale (ULSI) devices was also proposed [2]. Besides
these many applications, a-C films are materials with considerable interest from an intrinsically basic point of view.
Their properties are essentially controlled by the ratio between the number of carbon atoms in sp2 and sp3
hybridization states. The amorphous atomic network is usually described by sp2 carbon nanoclusters bonded to each
other by sp3 bonds. The properties of the films, which are closely related to their microstructure, can be tuned by the
deposition technique employed and by the growth conditions, with the energy of the impinging species playing the
main role in the control of the carbon bonding hybridization [1].
In the last few years there has been a strong interest in the study of nitrogenated amorphous carbon films [3]. One
of the main reasons of this research effort was the intent to synthesize the β-C3N4 solid, proposed by Liu and Cohen
to have mechanical properties comparable to that of diamond [4]. Despite this, no clear experimental evidence of the
formation of β-C3N4 has been presented until now [3]. An important fraction of this research effort was dedicated to
the study of hydrogenated amorphous carbon-nitrogen films (a-C(N):H) deposited by plasma enhanced chemical
vapor deposition, PECVD. Nitrogen incorporation in a-C:H films was found to modify the structure and the
mechanical properties [5-7] of these films, as well as their electrical and optical properties [8]. It results in a strong
decrease of the fraction of carbon atoms in sp3 hybridization state [6]. In which concerns to the mechanical
properties, a strong reduction on the internal stress was observed, with minor changes in the mechanical hardness [5].
Concerning the modification of electrical and optical properties, it was found that nitrogen could electronically dope
a-C:H films, with the simultaneous reduction of the electronic defect density [8]. This makes possible the use of
a-C(N):H films as a semiconductor material [8]. It is clear that these so many applications are critically dependent on
the film surface properties.
One of the ways to modify surface properties in a controlled way is by low-energy ion bombardment that can be
achieved by plasma treatment. For example, argon plasma is used to increase the surface roughness of a-C(N):H
films thereby creating a more efficient electron emitter [9]. However, there are only a few studies in the surface
nitrogen incorporation in a-C:H films by nitrogen plasma processing [10,11]. In this work, we will review some of
the effects of the nitrogen incorporation into a-C:H films during film growth and by post-deposition N2 rf-plasma
treatment.

CP669, Plasma Physics: 11th International Congress on Plasma Physics: ICPP 2002
edited by I. S. Falconer, R. L. Dewar, and J. Khachan
© 2003 American Institute of Physics 0-7354-0133-0/03/$20.00
354
NITROGEN INCORPORATION DURING FILM GROWTH
The incorporation of nitrogen in a-C:H films deposited by PECVD was studied for several precursor gas-
mixtures involving different hydrocarbon gases [3,6]. In the case of methane, the use of ammonia or N2 as a source
of nitrogen results in a-C(N):H films with properties governed by the amount of nitrogen incorporated in the
amorphous network. In Table 1 we present results obtained from films deposited by PECVD using CH4-N2 mixtures
with self-bias voltage (Vb) fixed at –350V and total pressure of 8Pa. It is clear that the incorporation of N occurs at
the expenses of the carbon content of the films. The film density remains nearly constant at 2.2 ± 0.2 g/cm3. The
observed deposition rate reduction was attributed to the onset of a kind of chemical sputtering process arising from
energetic N2+ bombardment of the film growing surface [12]. However, there is no justification for the absence of
the symmetric situation in carbon-nitrogen film deposition process, i.e., nitrogen atoms being sputtered out by
carbon carrying ions. Todorov at al. [13] considered both processes in the simulation of ion beam deposition of
carbon-nitrogen films. In addition, they included the possibility of N2 evaporation during film growth. This situation
is likely to occur since N-atom subsurface penetration must favor N-N bond formation with N atoms already in film.
The formation of other volatile species, as CN, can also occur. These mechanisms are increasingly more probable
for deposition conditions that result in increasing nitrogen contents, reducing the deposition rates. Therefore, the
growth kinetics of carbon-nitrogen films may be pictured as a competition between aggregation and erosion
resulting from the impinging of different species on the film-growing surface.

TABLE 1. Film composition and deposition rates as functions of the nitrogen partial pressure in CH4-N2 atmospheres.
N2 partial pressure (%) Composition (at.%) Deposition rate (nm/s)
C N H
0.0 86 0 14 12.5
12.5 79 4 16 12.5
25.0 79 8 14 11.0
37.5 76 9 15 6.0
50.0 72 11 17 2.5

In which concerns the structural modifications, the incorporation of nitrogen increases the size or the number of
sp2-carbon clusters, as is clear from Raman and electron energy loss (EELS) results. In fig. 1a, we show the ratio
between the intensities of the D- and G-bands, the main features of the Raman spectra obtained from a-C(N):H
films. This increase is interpreted as being due to an increase in size or in number of the sp2-clusters [6]. The
increase of the fraction of sp2-hybridized carbon atoms measured by EELS (fig. 1b) supports this interpretation.

1.5 100
(a) (b)

1.2
80
sp fraction
ID/IG

0.9
2

60

0.6
40
0 3 6 9 12 0 3 6 9 12
Nitrogen content (at.%) Nitrogen content (at.%)

FIGURE 1. (a) ID/IG intensity ratio and (b) sp2 fraction as functions of the incorporated nitrogen. The lines are eye-guides only.

It is well known that the high internal compressive stress observed in a-C:H films may be viewed as a result of
the material overconstraining. This mean that the relatively high carbon sp3 fraction observed in hard a-C:H films
causes the mean atomic coordination number to be higher than the ideal value predicted for a fully constrained
network [14]. In this scheme, any stress relief process may be strongly coupled to a reduction of the coordination
number. In the case of a-C(N):H films, the important decrease in the internal compressive stress upon nitrogen
incorporation (see fig. 2a) is conceived as a combination of the chemical composition and hybridization states.

355
Besides the nitrogen incorporation itself, the observed increase in the fraction of carbon atoms in sp2-hybridization
state is a source of decrease in the mean coordination number. Thus, for 8.4 at.% of nitrogen content in a-C(N):H
films, the hardness shows only a slight reduction (fig. 2b) while the internal stress is reduced by a factor of two.

3.0 27
Internal stress (GPa) (a) (b)
24

Hardness (GPa)
2.5

21
2.0

18
1.5

15
0 2 4 6 8 0 2 4 6 8

Nitrogen content (at.%) Nitrogen content (at.%)

FIGURE 2. (a) Compressive internal stress and (b) Hardness as functions of the incorporated nitrogen. The lines are guide-eyes.

NITROGEN PLASMA SURFACE PROCESSING


The effects of the N2 rf-plasma treatment of a-C:H films deposited by PECVD (precursor gas: CH4, P= 8Pa, Vb=
-350V) were investigated by a multitechnique approach. The total pressure in the plasma chamber was 3Pa and Vb
was in the range between -50 and -500V. At this pressure, ions mean free path is of the order of the plasma sheath.
The films were not exposed to air before submitted to plasma processing. No kind of sequential treatment was
performed. The results of erosion rate, friction coefficient and amount of nitrogen incorporated as functions of the
treatment time indicate that, for these plasma conditions, a steady state regime is achieved within tens of seconds.
In fig. 3a we present the erosion rate as a function of Vb. The increase of the erosion rate with self-bias voltage
may be at least partially ascribed to the increased N2+ ion current, since higher Vb is achieved by increasing the rf
power fed to the substrate electrode (from 4 to 55W, in the range of Vb studied). On the other hand, the self-bias
increase also results in the increase of the N2+ energy, and this may also affect the erosion process since both
chemical and physical sputtering of carbon films by N2+ increases with the ion energy [15]. In fig. 3b, we show the
depth profiles obtained by medium ion scattering spectrometry (MEIS) from a sample treated 20 minutes at -500V.
The total amount of nitrogen at the treated surface is nearly the same (~5x1015 atoms/cm2) and independent of the
self-bias. Probably, the increase of the sputtering yield with the ion energy compensates the increase of ion flux due
the increase of gas dissociation when higher power was applied to the plasma to achieve higher values of Vb. The
chemical bonds were probed by x-ray photoelectron spectroscopy and, as in the case of a-C(N):H films deposited by
PECVD, we could identify two chemical environments, one that corresponds to N substitutional in a graphitic-like
configuration and the second one that corresponds to N single bonded to C (sp2) or to C≡N bonds [16].

9 Plasma conditions: (b)


0.9
erosion rate (nm/min)

∆t = 20 min
Composition (at.%)

P = 3Pa
6 0.6 C
N
O
3 0.3

(a)
0.0
0 0 20 40 60 80
0 200 400 600
Vb (-V) Depth (nm)

FIGURE 3. (a) Erosion rate as function of the self-bias and (b) depth profiles obtained by MEIS from an a-C:H films
treated by N2 plasma during 20 minutes with Vb= -500V and pressure of 3Pa. The depth profiles are quite similar for all Vb.

356
The plasma treatment modifies the film surface properties, as is illustrated in fig. 4. A progressive increase of the
friction coefficient upon the increase of Vb was measured by atomic force microscopy. Films submitted to plasma
treatment with Vb higher than -100V are slightly rougher when compared with as-deposited a-C:H films. However, a
factor of three higher roughness values was determined in films treated at Vb = -100V. In our plasma conditions, the
energies of the N2+ that impinges the film are controlled essentially by the Vb. In these cases, N2+ ions break when
hit the film surface and each fragment carries half of the incident energy. It is important to note that the displacement
energies of graphite and diamond are 25 and 80eV, respectively. Thus, at low N2+ energies and depending on the
sp2/sp3 ratio, subimplantation cannot occur. In this case, surface diffusion tends to generate ordered clusters of high
sp2 content with structures closer to the thermodynamically stable graphite phase. Such clusters that can be nucleated
also by post-deposition annealing lead to high surface roughness [17]. However, more experimental investigations
are needed in order to understand the friction coefficient behavior of plasma processed films. In fact, no direct
correlation between nitrogen content and surface roughness with friction coefficient could be determined.

(a) 0.24 (b)


0.9
Roughness (nm)

Friction coeficient
0.22
0.6
0.20

0.3
0.18

0 150 300 450 0 150 300 450


Vb (-V) Vb (-V)

FIGURE 4. (a) RMS surface roughness and (b) friction coefficient as functions of Vb. The lines are to guide the eyes only.

ACKNOWLEDGMENTS
This work is partially supported by the Brazilian agencies: CNPq and FAPERJ.

REFERENCES
1. Robertson, J., Mater. Sci. Eng. R 37, 129 (2002).
2. Grill, A., Diamond Rel. Mater. 10, 234 (2001).
3. Muhl, S. and Mendez, J. M. , Diamond Rel. Mater. 8, 1809 (1999).
4. A. Y. Liu and M. L. Cohen, Science 245, 841 (1989).
5. Franceschini, D. F., Achete, C. A. and Freire Jr., F. L., Appl. Phys. Lett. 60, 3229 (1992).
6. Freire Jr., F. L., Jpn. J. Appl. Phys. 36, 4892 (1997).
7. Freire Jr., F. L., J. Non Cryst. Solids 304, 251 (2002).
8. Silva, S. R. P., Robertson, J., Amaratunga, G. A. J., Raferty, B., Brown, L. M., Schwan, J., Franceschini D. F. and Mariotto,
G., J. Appl. Phys. 81, 2626 (1997).
9. Liu, X. W.,Tsai, S. H., Lee, L. H., Yang, M. X., Yang, A. C. M., Lin, I. N. and Shih, H. C., J. Vac. Sci. Technol. B 18, 1840
(2000).
10. Hong, J., Granier, A., Goullet, A. and G. Turban, Diamond Relat. Mater. 9, 573 (2000).
11. Castañeda, S. I., Espinoza, V. A. A., Freire Jr., F. L., Franceschini, D. F. and Jacobsohn, L. G., Nucl. Instr. Meth. B 175-177,
699 (2001).
12. Clay, K. J., Speakman, S. P., Amaratunga, G. A. J., Silva and S. R. P., J. Appl. Phys. 79, 7227 (1996).
13. Todorov, S., Marton, D., Boyd, K. J., Al-Bayati, A. H. and Rabelais, J. W., J. Vac. Sci. Technol. A 12, 3192 (1994).
14. Angus, J. C. and Jansen, F., J. Vac. Sci. Technol A 6, 1778 (1988).
15. P. Hammer, P. and Gissler W., Diamond Relat. Mater. 5, 1152 (1996).
16. Ripalda, J. M., Díaz, N., Román, E., Galán, L., Montero, I., Goldoni, A., Baraldi, A., Lizzit, S., Comelli, G. and Paolucci, G.,
Phys. Rev. Lett. 85, 2132 (2000).
17. Peng, X. L., Barber, Z. H. and Clyne, T. W., Surf. Coat. Technol. 138, 23 (2001).

357

You might also like