Download as pdf or txt
Download as pdf or txt
You are on page 1of 161

Progress in Materials Science 47 (2002) 1±161

www.elsevier.com/locate/pmatsci

Laser nitriding of metals


Peter Schaaf
Zweites Physikalisches Institut der UniversitaÈt GoÈttingen, Bunsenstrasse 7/9,
D-37073 GoÈttingen, Germany

Received 3 January 2000; accepted 27 January 2000

Abstract
Laser nitriding can be described as the irradiation of metal surfaces by short laser pulses in
nitrogen containing atmospheres. This may lead to a strong take-up of nitrogen into the metal
and nitride formation which can improve the metal's surface properties, e.g. the hardness or
the corrosion and wear resistance. Here, the laser nitriding of iron, carbon steel, stainless steel,
and aluminum was investigated employing a combination of complementary methods. Ion
beam analysis (Rutherford Backscattering Spectroscopy and Resonant Nuclear Reaction
Analysis) was employed for element and isotope pro®ling. MoÈssbauer spectroscopy and X-ray
di€raction were used for phase analysis. Surface pro®lometry, optical and electron micro-
scopy revealed the surface topography and morphology obtained after laser nitriding.
Microhardness measurements by the nanoindentation technique characterized the mechanical
surface properties obtained by the treatment. By this combination of methods it became pos-
sible to resolve the in¯uence of the treatment parameters (laser ¯uence, number of pulses, spot
size, spatial intensity distribution, and gas pressure) in di€erent materials treated (iron, carbon
steels and stainless steel). It is shown that laser nitriding is a complex process, composed of
several superimposed e€ects. Laser heating, melting and evaporation in combination with plasma
formation and the generation of laser-supported absorption waves are the essentials of the
process. Pressure- and plasma-enhanced dissolution and di€usion of nitrogen in combination with
macroscopic material transport (piston e€ect, convection, fall-out) are further important e€ects
determining the results. Additional marker experiments and laser treatments in isotopically
enriched nitrogen atmospheres allowed to analyze these e€ects and to develop scenarios for
the nitriding process and the material transport mechanisms. A simulation of the nitrogen
depth pro®les for the single spot irradiations was derived, whose results are in good agreement
with the experimentally observed pro®les. # 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Nitriding; Nitri®cation; Short laser pulses; Laser material interaction; Laser generated plasma;
Laser absorption waves; Nitride formation; Nitrogen di€usion; Convection; Surface melting; Quenching;
Ion beam analysis; Rutherford backscattering spectrometry; Resonant nuclear reaction analysis; MoÈss-
bauer spectroscopy; Nanoindentation hardness measurements

E-mail address: pschaaf@uni- goettingen.de (Peter Schaaf).

0079-6425/02/$ - see front matter # 2002 Elsevier Science Ltd. All rights reserved.
PII: S0079-6425(00)00003-7
2 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Contents

1. Introduction...................................................................................................... 4

2. The Fe±N system .............................................................................................. 5


2.1. The Fe±N phase diagram........................................................................ 6
2.2. Solubility of nitrogen .............................................................................. 8
2.3. Lattice constants and densities ............................................................... 9
2.4. Magnetic properties .............................................................................. 11
2.5. Di€usion constants ............................................................................... 12

3. Nitriding processes.......................................................................................... 12
3.1. Gas nitriding ......................................................................................... 14
3.2. Salt bath nitriding ................................................................................. 17
3.3. Plasma nitriding .................................................................................... 18
3.4. Nitrogen implantation .......................................................................... 18
3.5. Plasma immersion ion implantation ..................................................... 19

4. Laser±material interactions............................................................................. 19
4.1. Laser ..................................................................................................... 20
4.2. Laser±material interactions................................................................... 20
4.3. Temperature pro®les ............................................................................. 23
4.4. Laser supported absorption waves........................................................ 29
4.5. Material transport phenomena ............................................................. 33

5. Experimental methods .................................................................................... 36


5.1. Materials and sample preparation ........................................................ 36
5.2. Laser treatments ................................................................................... 37
5.3. Ion beam analyses................................................................................. 41
5.3.1. Rutherford backscattering spectrometry (RBS) ........................ 41
5.3.2. Resonant nuclear reaction analysis (RNRA) ............................ 46
5.4. MoÈssbauer spectroscopy ....................................................................... 54
5.5. Microhardness by nanoindentation measurements............................... 59
5.6. Surface pro®ling.................................................................................... 61
5.7. Plasma imaging ..................................................................................... 61
5.8. Further analyzing methods ................................................................... 62

6. Raw beam irradiations.................................................................................... 62


6.1. Surface pro®les ..................................................................................... 63
6.2. Lateral nitrogen pro®les........................................................................ 69
6.3. Nitrogen depth pro®les ......................................................................... 70
6.4. Irradiation of marker layers.................................................................. 73
6.5. Isotopic experiments and modeling of depth pro®les ........................... 74
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 3

6.6. In¯uence of the energy density ............................................................. 80


6.7. In¯uence of the spot size....................................................................... 82
6.8. Microhardness measurements ............................................................... 86
6.9. Summary of raw beam irradiation........................................................ 90

7. Homogenized beam irradiations ..................................................................... 91


7.1. Surface pro®les ..................................................................................... 92
7.2. Lateral nitrogen pro®les........................................................................ 94
7.3. Nitrogen depth pro®les ......................................................................... 96
7.4. Irradiation of marker layers.................................................................. 97
7.5. Isotopic experiments ............................................................................. 99
7.6. Modeling of nitrogen depth pro®les ....................................................101
7.7. In¯uence of the energy density ............................................................103
7.8. Microhardness measurements ..............................................................105
7.9. Summary of the homogenized beam irradiation..................................106

8. Phase formation.............................................................................................107
8.1. Overview on MoÈssbauer data of the Fe±N system ..............................107
8.2. Austenite ..............................................................................................108
8.3. Epsilon-nitride .....................................................................................109
8.4. Results for laser nitrided iron ..............................................................111

9. In¯uence of the nitrogen gas pressure ...........................................................117


9.1. Surface morphology.............................................................................117
9.2. Nitrogen concentrations ......................................................................119
9.3. Phase formation...................................................................................125
9.4. Microhardness .....................................................................................126
9.5. Summary of pressure dependence ........................................................128

10. In¯uence of alloying elements........................................................................128


10.1. In¯uence of the carbon content ...........................................................128
10.2. Laser nitriding of stainless steel ...........................................................135
10.3. Comparison with laser nitriding of aluminum.....................................141
10.4. Summary of alloys ...............................................................................142

11. Summary and conclusions .............................................................................142


11.1. Summary of raw and homogenized beam irradiations ........................142
11.2. Conclusions and outlook .....................................................................144

Acknowledgements................................................................................................147

References .............................................................................................................148
4 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

1. Introduction

Nitriding is a common method for improving the hardness, mechanical properties,


wear and corrosion resistance of metals [1,2]. Many machine parts, work pieces and
tools (toothwheels, camshafts, cylinder liners, rocker arms, bio-implants) are
industrially nitrided in order to improve their tribological and chemical properties.
In recent years, more and more industrial applications of laser-based processes have
been developed. As examples, laser cutting [3±5], laser welding [6±8], laser hardening
[9±15], laser remelting [16±18], laser alloying [19±21], laser cladding [22±24] and laser
glazing [25±27] should be mentioned. Further applications are laser drilling, laser
writing and laser structuring [28]. Applications involving laser-induced chemical
reactions in particular, are becoming more and more popular [28]. There is also
large interest in the pulsed laser deposition (PLD) of metallic layers [29±32] or dia-
mond-like carbon [33]. The PLD can also be carried out in reactive atmospheres
(YBaCuO [34±28], TiN [36]). Chen and coworkers [37] investigated the interaction
of ambient background gas with the laser plume. Troe et al. [38] published a detailed
study on the pulsed laser ablation of polymers (spot size, pulse length, plasma
attenuation). Sub-picosecond ablation of metals [39] or laser pulse ablation of
liquids were also investigated [40].
Meanwhile, it is well established that the irradiation of iron and other metals with
pulses of an excimer laser in a nitrogen atmosphere or air leads to a huge take-up of
nitrogen into the irradiated surfaces [41±57]. The laser nitriding e€ect has been
demonstrated for various materials and for di€erent laser systems [58±67], e.g. for Ti
[59], where even the formation of stoichiometric, adherent layers of TiN was repor-
ted [36]. Laser nitriding, when applied to iron [41±43,61±64], showed improved
hardness and corrosion resistance of the irradiated samples. The laser nitriding of
aluminum and AlSi alloys was investigated by Barnikel and coworkers [68±70],
while Meneau et al. [71] reported the formation of AlN by excimer laser irradiation
in nitrogen gas. Laude et al. [72] investigated the excimer laser irradiation of cera-
mics. In comparison to conventional nitriding methods, the use of short laser pulses
has several advantages. Due to the small heat-a€ected zone, both in depth and lat-
eral dimension, pieces sensitive to heat and of complex shape can be modi®ed. Laser
nitriding is a very ecient technique, allowing an accurate spatial control of the
surface treatment without any undesired heating of the substrate.
The basic mechanisms of the laser nitriding process, however, are still rather
poorly understood, as they are closely related to the laser-induced plasma above the
surface and to the complicated laser±plasma±material interactions. The present
report focuses on the excimer laser nitriding of iron. Because of the complexity of
the process, based on the combined laser±material, laser±plasma and laser±plasma±
material interactions (heating, convection, plasma formation, ¯uid dynamics,
plasma dynamics, di€usion and metallurgical processes)1 a combination of numer-
ous methods is required for the investigation of these interactions and for ®nally
optimizing an industrial production process.

1
An investigation of the laser nitriding is very close to the opening of Pandora's box [73].
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 5

In the present study, the analyzing methods were focused on the observation of
the changes induced in the sample after ®nishing the laser nitriding process. The
investigation of the plasma, the plasma dynamics and the laser-plasma interactions
were not accessible experimentally and are far beyond the scope of this work. The
most important aim was to gain insights into the in¯uence of the laser nitriding
parameters on the nitrogen take-up, the lateral and depth distribution of nitrogen,
the phase formation and on the resulting surface properties (hardness). Although
the laser nitriding of iron can also be accomplished in air, resulting in a nitriding
e€ect not hindered by a thin oxide layer of about 30 nm formed at the surface [42],
the present work is restricted to pure nitrogen atmospheres.
This report is organized in the following way: ®rst a brief description of the Fe±N
system (Section 2), a short overview of the established nitriding methods (Section 3)
and some basic details of laser material interactions (Section 4) will be given. After
that, the laser treatment and the measuring methods of ion beam analysis, MoÈss-
bauer spectroscopy, surface pro®lometry and microhardness measurements will be
explained (Section 5). In the Sections 6±9 a detailed analysis of the nitrogen pro®les,
laterally and in depth, and the phases formed will be presented, together with the
results of the surface pro®lometry and the microhardness. The dependence of
these quantities on the intensity pro®le (raw beam and homogenized beam) and
other treatment parameters (number of pulses, laser energy density, nitrogen gas
pressure) gives insight into important details of the irradiation parameters. With
additional experiments using isotopic enrichments of the nitriding atmosphere,
together with marker experiments, it was possible to model and simulate the devel-
opment of the nitrogen depth pro®les with the number of laser pulses. The simila-
rities and the di€erences observed after the raw-beam and homogeneous laser
irradiations are compared and included into the developed scenarios [48,55,74,75].
Finally, ®rst results on actually used materials (carbon steel, stainless steel) are
reported in Section 10.

2. The Fe±N system

The phase diagram of the Fe±N system is being reviewed. Additional important
data, e.g. densities, transition temperature, di€usion constants and solubilities, are
also summarized in this chapter.
Iron nitrides and other transition metal nitrides exhibit interesting physical, che-
mical, mechanical, electrical and/or magnetic properties and therefore deserve sci-
enti®c interest. Due to their physical properties (high hardness, high melting point),
transition metal nitrides, such as TiN and CrN, have been used in industrial appli-
cations for a long time [65,76±78]. The compounds and phase equilibria of the sys-
tem Fe±N play an important role in steel production and surface hardening [2,79±
81]. The cost intensive surface hardening should increase the hardness and enhance
the lifetime and resistance against wear and corrosion [2,81]. Besides these techno-
logical aspects, the iron nitrides have also attracted interest due to their magnetic
properties. The search for the giant magnetic moment of the 00 -phase is an example
6 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

for that [82±87], including measurements on single crystal 00 -Fe16N2 produced by


molecular beam epitaxy [87]. In the past few years also the high-performance cera-
mics such as nitrides of boron, aluminum and silicon [88] have attracted interest.
Based on their excellent thermal stability and heat conductivity, their use in inno-
vative technologies (engines, computer, nuclear fusion and environmental technol-
ogy) can be expected [68,69,89].

2.1. The Fe±N phase diagram

The system Fe±N has been the object of intense investigations for the past 150
years till date. The take-up of nitrogen into iron by a gas ¯ow of ammonia was
observed for the ®rst time in 1828 by Savart [90]. The structures of the nitrides
formed during these treatments were analyzed in the ®rst half of this century by
HaÈgg [91,92], and the thermodynamic mechanisms were treated empirically and
theoretically by Fry [93,94] and Lehrer [95,96]. The ®rst detailed analysis of the Fe±
N system was compiled by Jack [97±103]. He found the solid solutions martensite
and austenite, as well as the compounds 00 , 0 , " and . Based on these and other
results [95,96,104,105] Jack compiled in the ®fties a phase diagram which is basically
still valid. When comparing his phase diagram with the currently accepted one of
Wriedt, Gokcen and Nafziger [106] one hardly ®nds large di€erences, except for the
narrower existence regions of 0 and .
The Fe±N system consists of several interstitial solutions ( , , "), chemical com-
pounds ( 0 -Fe4N, -Fe2N) and metastable phases ( 0 -martensite, 00 -Fe16N2)
[106,107]. Fig. 1 shows the Fe±N phase diagram, which was extended by a calcula-
tion for the high temperature region [108] as above 1200 K no experimental data
exists. This calculation is based on the common CALPHAD formalism of Gibbs
free energies [109,110].

Fig. 1. Fe±N phase diagram (after Refs. [106,108]).


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 7

More recent calculations for the low-temperature region by Kooi, Somers and
Mittemeijer [111±113] have re®ned some phase boundaries and phase structures, but
con®rmed the basic principles of the phase diagram. These thermodynamic calcula-
tions have also been extended to ternary and even quaternary systems, which are
important for steel research, such as the Fe±N±C system [114], as well as the Fe±Cr±
N and the Fe±Cr±Mn±N systems [115±117]. The special features of the Fe±N system
are summarized in Table 1 based on Ref. [118].
Due to the practical importance and the relative simplicity of an only binary phase
diagram, considered to be most accurate in the version of Wriedt, Gokcen and
Nafziger [106,119], not too many questions should remain unsolved there. In fact,
the expressions used to describe the Gibbs free energy of Fe±N phases
[108,107,110,120±123] are based mainly on the Hillert±Sta€anson approach for
interstitial solid solutions [124,125]. This approach ignores the important in¯uence
of the short range order and long range order present in the Fe±N nitrides. Some
attempts have been made to include the long range order into the thermodynamical
description [112,113,126±130] and led to a major improvement of the theoretical
understanding of the Fe±N phase diagram. Nevertheless, the question of the ®nal
overall ordering of nitrogen atoms in the "-phase, for example, is still open.
The austenite g-Fe(N) is only stable above 873 K, but can be retained at room
temperature by fast cooling. The start temperature of its transformation into mar-
tensite and the extent of transformation (Ms , Mf ) depend on the nitrogen content
and the cooling rate [2,9,131,132]. If the nitrogen concentration is higher than cN ˆ
8:6 at.%, this transformation can be completely avoided and all austenite can be
retained at room temperature as a metastable phase [106]. The stability of other iron
nitrides below 625 K has also been investigated.
The " iron nitride has a hexagonal crystal structure (P312 or P6322) [102,133] and
nitrogen is soluble between 15 and 33 at.%. Increasing the nitrogen concentration
from 33 at.% to 33.2 at.% leads to an anisotropic distortion of the " nitride lattice

Table 1
Transition points in the Fe±N system

Reaction Compositions (at.%) Temperature (K) Reaction type

para $ ferro 0 1043 Curie point


$ 0 1185 Allotropic
$  0 1667 Allotropic
L $  0 1811 Melting
" $ 0 19.5 953 Congruent
" $  33.3 753 Congruenta
… † $ … † ‡ 0 8.8 0.4 19.3 865 Eutectoid
" $ … † ‡ 0 15.9 10.3 19.3 923 Eutectoid
… † $ … † ‡ 00     Probably eutectoidb
L ‡ …† $ … † 11 3.5 6 1768 Probably peritectica,c
L ‡ … † ‡ "     Peritectic or eutectica
a
Not observed, conceptionally possible.
b
Not observed, possibly metastable.
c
Composition estimated based on calculation [118].
8 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

and the orthorhombic -Fe2N line phase is formed [134]. For technical applications,
especially the hexagonal " is of interest, due to its high hardness of 7.1(8) GPa and
improved corrosion resistance [135,136]. The "-nitride with nitrogen contents below
27 at.% is metastable at room temperature. With decreasing nitrogen content it
decomposes into the " phase with a higher nitrogen concentration and the 0 -phase
[106,107]. Its decomposition into " and the 00 -phase has also been reported
[137,138]. There is an ordering of the nitrogen atoms in this nitride [102,139], but
what this ordering exactly looks like is still the topic of scienti®c dispute [133,140±
142]. This will be described in more detail below (Section 8.3).
In recent years an increasing number of publications about a new compound FeN
appeared [44,143±150], which has been prepared by sputtering methods. In the ®rst
structure determination a NaCl structure was found [144], which is known to be the
favored crystal structure of the transition metal nitrides of type MN (M = Sc, Ti, V,
Cr) [151]. Another powder di€raction analysis revealed a ZnS structure [146], which
is a strange tetrahedral coordination of iron. Nakagawa and coworkers [152] also
reported the production of a NaCl-type phase with a lattice constant a = 4.5 AÊ and
denoted it as 000 -FeNy, y  0:65. This phase is intermixed with another nitrogen-rich
phase 00 -FeNy with y  0:91 and ZnS structure (a = 4.33 AÊ). This mixture of ZnS
( 00 ) and NaCl ( 000 ) structure was also reported in subsequent publications [147±
152,153]. The phase is stable up to about 600 K where it decomposes into Fe2N
[146,147]. So the question of the ®nal structure and the stability region of this phase
remains open [149]. The crystallographic structures of all the phases in the Fe±N
system can now be summarized in Table 2 (from Refs. [106,107,146,148,150,152]).

2.2. Solubility of nitrogen

The solubility of nitrogen in a-Fe is very low, only up to a maximum of 0.4 at.%
at 865 K. The g-phase has a maximum solubility of 10.3 at.% at 923 K. The mar-
tensite 0 may contain up to 12.0 at.% nitrogen. The composition ranges of the 00 -
phase and the compounds Fe4N and Fe2N are very small. The interstitial solution of

Table 2
Structure and lattice constants of the phases in the Fe±N system

Phase Alias Structure Lattice constants cN (at.%)

a (AÊ) b (AÊ) c (AÊ)

-Fe Ferrite bcc (Im3m) 2.8664 0.0


0 -Fe(N) Martensite bct (Im3m) 2.861 2.936 2.8
2.848 3.120 9.5
00 -Fe16N2 00 bct (I4/mmm) 5.72 6.29 11.1
-Fe(N) Austenite fcc (Fm3m) 3.572 0.0
0 -Fe4N Fe4N fcc (CaTi03) 3.798 20.0
"-Fex N " hcp (P63/mmc) 2.529 4.107 0.0
-Fe2N Fe2N (Pbcn) 4.8429 5.5413 4.4373 33.3
00 -FeN 00 fcc(ZnS) 4.3
000 -FeN 000 fcc (NaCl) 4.5
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 9

nitrogen in the hexagonal " has the widest range from about 15 at.% at the Fe-rich
side to at least 33 at.% at the N-rich side, and maybe even more. This region is dif-
®cult to investigate as extremely high N2 fugacities are involved, especially at ele-
vated temperatures.
The solubility of nitrogen in liquid iron in equilibrium with nitrogen gas at pres-
sure p is given by Sieverts law and its temperature dependence is small [106,107].
Measurements at 1833 K [154] and up to 50 bar at 1873 K by Feichtinger [155] lead
to the following equation:
 p  536:2 K
log cN ˆ 1=2 log 4:966 : …1†
1 Pa T

At the melting point T ˆ 1811 K and atmospheric pressure p ˆ 1013 hPa we only
have a solubility of cN ˆ 0:17 at.%. But it is worth mentioning that the solubility of
nitrogen in liquid iron can be signi®cantly enhanced, e.g. by a factor of about 3 by
an arc discharge above the liquid surface [156].

2.3. Lattice constants and densities

The lattice constants of the interstitial solutions depend on the nitrogen content cN
[106,107]. The lattice constant in 0 was determined by Somers and coworkers [157].
For the , 0 , and " phases the following equations are given by Kunze [107]:

: a ˆ 2:8664 ‡ 0:79  cN …2†

0 : a ˆ 2:8664 0:18  yN c=a ˆ 1:000 ‡ 0:91  yN …3†

: a ˆ 3:572 ‡ 0:78  cN …4†

 : a ˆ 2:519 ‡ 0:50  yN c=a ˆ 1:633 0:05  yN …5†

where all lattice constants are given in AÊ, cN as atomic fraction, and yN is the atomic
ratio, yN ˆ cN =…1 cN †.
For the "-phase, Schaaf et al. [44] performed a new compilation based on the lit-
erature data summarized by Wriedt [106], which is displayed in Fig. 2 and they
obtained the following results:

a ˆ 2:525…9† ‡ 0:747…29†  cN
c ˆ 4:238…7† ‡ 0:565…22†  cN
c=a ˆ 1:680…3† 0:25…1†  cN : …6†

With the lattice constants also the densities of the nitride phases vary with the
nitrogen content, which is important for the ion beam analyses. Some experimental
data [140,158] are displayed in Fig. 3, together with the theoretical curves calculated
10 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

from the lattice constants (atoms per elementary cell divided by elementary cell
volume) as given in Eqs. (2), (4) and (6). The following approximation is used for the
atomic density as a function of the nitrogen concentration up to cN ˆ 0:33 (cN is
used here as atomic fraction):

4
1024  p
  3  … 1 cN †
N…cN † at:=cm3 ˆ …7†
27:02 ‡ 19:59  cN ‡ 4:496  c2N ‡ 0:3153  c3N

The g-phase always occurs intermixed with the , 0 -phase, so that in this region a
mean density value can be assumed. The density of the new FeN phase has not yet

Fig. 2. Dependence of the lattice constants a and c of the "-phase on the nitrogen concentration cN .

Fig. 3. Atomic densities in the Fe±N system (open squares measured, circle from structure).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 11

been determined experimentally and the theoretical value based on the preliminary
structure determinations [146±150] is given in Fig. 3. Obviously, the atomic density
can be well described by the theoretical density of the " phase up to cN ˆ 0:33.
Above, the density of FeN is taken as a constant value N ˆ 10:02  1022 at./cm3.

2.4. Magnetic properties

Also the magnetic properties of the phases in the Fe±N system have been
investigated [96] since the beginning of the studies in that system. The transition
temperatures of the various phases are summarized in Table 3. The magnetic prop-
erties of the 00 and 000 phases are not ®nally determined. Both are paramagnetic
at room temperature and ferro- and/or antiferromagnetic at lower temperatures
[148,146,153].
The Curie temperature TC of the "-nitride changes steadily with the nitrogen
concentration cN from 535 to 0 K [95,106] as shown in Fig. 4. At room tem-
perature the " phase is ferromagnetic for nitrogen concentrations of 17.5±30 at.%

Table 3
Magnetic transitions for Fe±N phases

Phase Transition temperature Ttr (K) Reference

TC ˆ 1043 K [131]
0 ; 00 Ferromagnetic, but decomposing below TC [106]
TN ˆ 67 K for cN ˆ 0 [159,160]
No TC for cN ˆ 7:4 9:9 at.% [161]
0 TC ˆ 752 775 K for (19.5±20.1 at.%) [96]
0 TC ˆ 753 781 K for (19.5±20.1 at.%) [106]
 TC ˆ 4 K at 33.2 at.% [160]
 TC < 4 K at 33.4 at.% [162]

Fig. 4. Curie temperature of the " phase (solid line according to Eq. (8), data from Ref. [106]).
12 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

and paramagnetic at 16.1±17.5 at.% and 30±33.3 at.% [163,160]. The variation of
the Curie temperature with the nitrogen concentration cN can be approximated via

T C ‰KŠ ˆ 11801 ‡ 1184  cN 108:5  c2N ‡ 2:83  c3N 0:028  c4N : …8†

high
The resulting values for TC = 300 K are clow
N ˆ 16:1 at.% and cN ˆ 30:6 at.%, in
good agreement with the values reported in literature [106,44] and also with the
values given above.

2.5. Di€usion constants

The di€usion constant D of nitrogen varies by several orders of magnitude [164±


167] and may also depend on the nitrogen concentration in the corresponding phase,
as determined [165±167] for various phases. The di€usivity of iron is hardly known
and is assumed to be much smaller than that of nitrogen. The temperature depen-
dence of D is given by

D ˆ D0 exp… Q=RT† …9†

where D0 is the di€usion constant for T ˆ 1 and Q the activation energy. As an


example the nitrogen di€usion in liquid iron is given by DL0 ˆ 2:86  10 3 cm2/s and
QL ˆ 61; 090 J/mol (1811±2023 K), whereas for the g-austenite the values are D…0 † ˆ
0:91 cm2/s and Q… † ˆ 168; 450 J/mol (1104±1667 K) [164] and for a-Fe D 0 ˆ 0:0042
cm2/s and Q… † ˆ 76; 120 J/mol (226±1184 K). The di€usion constant for nitrogen
di€usion in liquid iron at temperatures above 2023 K had to be extrapolated. For
the " phase no di€usion data are available because this phase decomposes at tem-
peratures above 800 K.

3. Nitriding processes

This chapter gives a short survey on the most common nitriding techniques. The
surfaces of steel are treated by a variety of methods for a variety of reasons.
Nitriding is a widely applied thermochemical surface treatment in which nitrogen is
introduced into steel or other iron-based alloys at elevated temperatures, typically
between 783 and 858 K [1,79]. The principal objectives for nitriding are: to obtain a
high surface hardness, improved wear resistance, enhanced fatigue life and better
corrosion resistance with generally a negligible change of the dimensions and prop-
erties of the workpiece itself.
A pure thermal surface treatment does not change the chemistry of the metal. The
surface is simply heated (austenized) and subsequently quenched. This surface
heating can be accomplished by a number of methods, e.g. by laser irradiation,
leading to laser hardening [9,15,168] or laser remelting [18]. If also the chemistry of
the surface is altered, we have a thermochemical surface treatment, for example laser
cladding and laser alloying [169±171].
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 13

One of the oldest and most important methods of thermochemical surface treat-
ment is the nitriding of steel [81], altering the chemistry of its surface layer. Nitrogen
can be added by bringing the surface into contact with an appropriate nitrogen-
containing environment at an appropriate temperature. As examples gas nitriding,
salt bath nitriding, plasma nitriding, pulsed laser deposition, reactive magnetron
sputtering and nitrogen implantation and plasma immersion ion implantation [172±
178,148] should be mentioned. Nitrogen is a relatively small atom and dissolves
interstitially in iron. The solubility is higher in austenite than in ferrite. As in heat
treatments, quenching from the austenite range will produce a hard martensite [9]. If
the nitriding is performed below the eutectoid temperature A1, which is the common
way, nitriding does not result in a hard martensite, but in nitrides which impart high
hardness and high wear resistance to the surface. These iron nitrides (and other
transition metal nitrides) exhibit interesting chemical, mechanical, electrical and/or
magnetic properties and therefore deserve scienti®c interest. Recently, also the ben-
e®cial nitriding above the eutectoid temperature has been reported [179±181]. If
nitriding is followed by an oxidation treatment, the atmospheric corrosion resistance
can increase signi®cantly [182,183].
Fig. 5 shows a calculated p±T diagram for the Fe±N system [123]. The nitrogen
activity is expressed here as the pressure of the molecular nitrogen gas. Due to the
low activity of molecular nitrogen, the iron-nitrides only form at very high nitrogen
gas pressures. It is seen, that pure N2 gas at atmospheric pressure does not dissociate
even if the surface layer is in the austenite range. While some metal nitrides can form
by N2 gas reactions at atmospheric pressure (VN, TiN, AlN), others form only in
contact with ammonia (NH3) [184], the phase -Fe2N even only by ¯owing pure
ammonia [97,139,185]. Thus, a peculiar property of the iron nitrides in relation to
their practical use is their metastability with respect to decomposition into N2 gas
and a-Fe. Under normal conditions (T < 673 K) however, the kinetic of the
decomposition is so slow that no signi®cant changes can be observed over the period

Fig. 5. T±p diagram for nitrogen gas (N2) and iron.


14 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

of application. At higher temperatures, where the nitrides are produced, this


decomposition cannot always be prevented, resulting in the coalescence of N2 ®lled
voids (precipitates) within the nitrided layer. Prolonged nitriding may thus give rise
to the formation of pores, channels or blisters inside the material with direct contact
to the nitriding environment [186±188].
This metastability of nitrides against N2 and a-Fe implies, that the production
does not occur in normal N2 gas. For example the virtual partial pressure of Fe3N at
773 K in equilibrium with N2 would be in the order of 5  1010 Pa (see Fig. 5). To
provide reactive nitrogen, i.e. a high nitrogen fugacity to the surface, several meth-
ods were developed. The most common and industrially used process is to use the
chemical equilibrium NH3 ) * 1=2N2 ‡ 3=2H2 , a mixture of ammonia with gaseous
nitrogen and hydrogen. Instead of NH3/H2 gas mixtures, also salt baths (containing
cyanides and cyanates) or ion implantation [189] deliver very active nitrogen and the
nitriding can be performed very easily [79]. The molecular nitrogen can also be
activated by a glow discharge in N2/H2 mixtures (plasma) above the surface [190].
Even Molecular Beam Epitaxy has been used [191] to produce iron nitrides. A short
sketch about the di€erent methods in industrial use is given below.
The formation of nitrided surface layers during nitriding is a complicated, com-
bined process, controlled by dissociation, absorption, desorption, di€usion, phase
formation and related processes, the temperature, ambient atmosphere (including
pressure, composition, temperature and possible activation by plasma) and the
substrate are the determining parameters. The nitrogen gradient is controlled by the
di€usion of nitrogen through the surface. This is valid only if the nitrogen is
deposited or absorbed faster than it di€uses inwards. Then, it can be assumed that
the nitrogen content at the surface is increasing, until the value is reached which
correspond to the thermodynamic equilibrium with the nitriding atmosphere. This
may be regarded as the saturation value, or surface value cs .
The di€usion equation can be solved with the appropriate boundary conditions,
e.g. the nitrogen concentration at the surface is ®xed at cN …0† ˆ cs and for samples
with large thickness d the nitrogen content at cN …z ˆ d  1† ˆ c0 , both of them
being independent of time. The solution of this di€usion problem is [192]:
 p
cN …z; t† ˆ c0 ‡ …cs c0 †  erf z= 4Dt : …10†

The temperature dependence of the di€usivity D was already given by Eq. (9) and its
values can be found for the various Fe±N phases in literature [164,165] (see Section 2.5).
Also Kirchheim and coworkers [193±195] measured the di€usion of nitrogen in metals.

3.1. Gas nitriding

Gas nitriding together with carbonitriding or nitrocarburization, if also a carbon


delivering medium is added, are the most used industrial processes. Millions of tons
per year [196] are treated this way and many e€orts are spend for their optimization
[197,198]. Although the thermodynamics of the reactions are well described, the
exact mechanisms during gas nitriding are still not fully understood. The practical
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 15

control of the process is still based on the empirical and theoretical results developed
in the thirties [93,95].
Nitriding and nitrocarburizing are the most versatile surface treatments for steels:
the fatigue, wear and corrosion properties can be improved signi®cantly. The high-
cycle fatigue resistance is improved by the di€usion zone consisting of a dispersion
of iron and alloying element nitrides in a ferrite matrix. Wear and corrosion prop-
erties are improved by the compound layer (white layer), composed of " nitride, on
top of this di€usion zone.
As mentioned before, gas nitriding uses ammonia or ammonia/hydrogen mixtures
to enhance the nitrogen activity. Ammonia easily dissociates into gaseous nitrogen
and hydrogen according to the chemical equilibrium

NH3 )* 1 N2 ‡ 3 H2 …11†
2 2

The equilibrium condition with the molar Gibbs free energies G is

1 3
G ˆ GNH3 GN GH ˆ 0; …12†
2 2 2 2

where G is given for a gas (nitrogen) by

GN2 ˆ G0N2 ‡ RT ln pN2 …13†

Inserting the standard Gibbs free energy G0 and the partial pressures pi of each gas
component the following Eq. (14) is obtained:
!
0 0 0 0 pNH3
G ˆ GNH3 1=2GN2 3=2GH2 ˆ RT ln 1=2 3=2 …14†
pN2 pH2

and ®nally for the nitrogen activity the following expression is derived:

1 pNH
RT ln pN2 ˆ G0 ‡ RT ln 3=23 …15†
2 pH2

The empirical equation

G0 ‰J=molŠ ˆ 45403 ‡ 128:94  T ‡ 10:785  T  ln…T† …16†

derived from tabulated values of Barin and Knacke [199], is valid between 725 and
973 K. The pressures are expressed in (Pa). Thus, the so called nitriding potential is
characterized by rN ˆ pNH3 =p3=2H2 , where normally the value of ln rN is mentioned.
For historical reasons, also the `Nitrierkennzahl' KN ˆ pNH3 =p3=2 H2 is used, where the
pressures are given in (bar). Its unit (bar 1=2 ) is usually omitted (see also Ref. [200]).
16 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

The `Lehrer diagram' presented in Fig. 6 can be derived from the thermo-
dynamical phase stability regions (see Fig. 5) and published thermodynamic data,
and by using Eq. (16) to transform the nitrogen activities into nitriding potential
[95,107,112,111,114]. It is seen, that the phase cannot be obtained below 900 K
and that the 0 is restricted to a narrow band. Thus, the ammonia mixture allows the
formation of the "-nitride at ambient pressure (1013 Pa) already at T = 773 K
starting with a fraction of 62 vol% NH3, i.e. a nitrogen activity KN ˆ 4 bar 1/2 or
ln rN ˆ 4:4, rN in Pa 1/2.
On the basis of the well known thermodynamics of the Fe±N system, the devel-
oping phases during nitriding can be predicted [107,112±114,141,201]. However,
thermodynamics is only describing the equilibrium state for the given nitriding
conditions. The actual constitution and composition of the nitrided layer and its
depth are determined by the kinetics [202]. The basic nitriding mechanisms are illu-
strated in Fig. 7.
Ammonia dissociates at the metal surface via a catalytic reaction into its elements,
according to the equilibrium reaction. By adsorption, di€usion and reaction, iron
nitrides are formed with decreasing nitrogen contents towards greater depths
according to the local nitriding potential. The compound layer is mainly composed
of 0 and " nitrides [167], because after the primary nucleation of 0 , the nucleation
of " on top of this 0 follows. Then after isolating the substrate from the nitriding
atmosphere, there is the nitrogen di€usion controlled growth of this "= 0 compound
layer [167]. The di€usivity of nitrogen in the various nitride phases has been deter-
mined [165,166]. This is important because the surface is constituted of a layered
structure of the phases "= 0 = . Therefore, there is not only a simple di€usion, but
one has to take into account the di€erent phases with their proper di€usion coe-
cients which are in addition concentration dependent. In fact, simple models can
solve this di€usion problem quite accurately [167,203±206]. Fitting the nitriding case
growth at T = 823 K by such a model with e€ective di€usion coecients and taking

Fig. 6. `Lehrer' diagram for gas-nitriding in ammonia.


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 17

into account the dependence of the di€usivity on the nitrogen content and on the
phase [167], led to hD"N i ˆ 7  10 15 m2 s 1. The value is in fair agreement with the
result given above. However, the total parabolic layer growth, which is typical for a
simple di€usion controlled process, has also to include the adsorption and deso-
rption rates at the surface [203,204,207,208]. The real layer growth is slower than
that of pure nitrogen di€usion.
On prolonged nitriding, pores may evolve, due to the metastability of iron nitrides
with respect to gaseous nitrogen and ferrite. Since the piece is normally at higher
temperature the nitrides are tending to decompose and the high partial pressures of
nitrogen may lead to the formation of pores [186,209]. Due to the volume changes
(see Fig. 3) also stresses are evolving in nitrided layers (k  800 MPa [210]), which
may lead to undesired cracks.
Also the bene®cial in¯uence of oxygen added during gas nitriding has been
demonstrated [211], although the nitriding potential has to be carefully controlled
[212±215]. There is a change in the structure of the nitrided surface as observed by
Somers and coworkers [203,216]. Especially for high alloy steels this is the only way
for a successful nitriding treatment [217]. Internal nitriding and external oxidation
proceed simultaneously in the early stages [211,218].

3.2. Salt bath nitriding

Salt bath nitriding is quite similar to the gas nitriding in ammonia mixtures, except
that the nitrogen activity comes from liquid salt baths containing the reactive
nitrogen (cyanides, cyanates). The thermodynamics and kinetics are analogous to
the gas nitriding process [203,204,219].

Fig. 7. Schematic view of the gas nitriding process.


18 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

3.3. Plasma nitriding

Plasma nitriding is a method of surface modi®cation using a glow discharge tech-


nology to introduce nitrogen into the surface of a metal which subsequently di€uses
into the material [220]. Due to its improved capabilities, this method becomes
increasingly popular. There, the sample and its surface is charged negative at a few
thousand volts and the nitrogen gas or the gas mixture is supplied at low pressures
(10 2±100 Pa). In the glow discharge and electrical potential gradient the gas
becomes ionized and the nitrogen ions are accelerated towards the negatively
charged surface. The surface of the heated piece is additionally heated by the
plasma. Advantages are the uniform surface treatment independent of the geometry
and even complicated pieces can be treated within reasonable times, say a couple of
hours [221]. Large vacuum chambers having plasma, pressure and temperature
controls are necessary for that. The increasing interest in the plasma surface treat-
ment is manifested in the growing number of conferences on this topic [221±226].
During plasma nitriding a hard case at the surface of the material is formed. Its
structure usually consists of a di€usion zone and a compound layer. Due to the ion
bombardment via plasma and potential di€erence, both the nitrogen di€usion
mechanisms and the parabolic law of layer growth are changed [227±230]. The
deviation from the parabolic growth has been explained by sputtering e€ects.
Dimitrov has developed a general di€usion model for surface plasma treatment
which takes into account the erosion of the material surface [205,231,206] based on
the original work of Wagner [232,233]. They observed and modeled an initial para-
bolic growth of the compound layer which then saturates after a certain plasma
nitriding time. At 823 K, parabolic growth occurs during the ®rst 42 min, then after
about 6 h nitriding saturates at dN ˆ 14 mm.
Also reactive magnetron sputtering may be assigned to this category of plasma
nitriding, since it also enhances the nitrogen activity by a magnetically con®ned plasma
to form nitrides, which then can be deposited onto almost any substrate. AlN, TiN,
NbN and also CrN, Cr2N [234,235,78] coatings with de®ned composition and
structure can be produced. Even the more dicult iron-nitrides can be accurately
deposited with desired composition and structure [44,178]. The nitrogen activity
could be increased here to unrivaled values, so that it was even possible to produce the
so far unknown phase FeN ( 00 , 000 ) by magnetron sputtering [44,145,147,148,152].
Soon after the development of the pulsed laser deposition (PLD) [29], also the
reactive deposition of metal nitrides using this technique with an appropriate (low)
nitrogen pressure in the irradiation and deposition chamber was investigated. The
laser pulse sputtering of AlN [236] and the pulsed laser deposition of TiN [237] are
examples for that. PLD also involves a plasma formation [29] and therefore also
may count to the category of plasma nitriding.

3.4. Nitrogen implantation

Another technique of changing the chemistry at or near the surface is by bom-


bardment with elemental ions or even molecules of sucient energy, so that they can
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 19

penetrate the surface and become embedded. This is called ion implantation
[188,238±252]. It allows to control the chemistry and structure, i.e. even amorphous
layers can be formed. The implantation requires an implanter delivering enough
acceleration voltage and a suitable quantity of ions to be retained in the surface. The
in¯uenced range, i.e. the nitriding depth, is limited by the range of the ions. It is
typically below the micron range, e.g. 1.4 mm for 3 MeV N+ into Fe (see [253]), if
the implantation is carried out at modest temperatures (<500 K). Di€usion can be
enhanced by higher temperatures or by radiation enhanced di€usion [254±256].
The use of low-energy, high-current-density nitrogen ion beams for surface mod-
i®cation of stainless steel led to dramatic improvements in tribological behavior
[250,257,248] as well as enhanced corrosion resistance [258,259]. The high current
densities lead to much thicker layers for the same processing time as compared to
normal ion beam implantation, plasma nitriding, and gas nitriding performed at the
same temperature of 673 K [248,249]. The formation of several mm thick, highly
nitrogen containing ®lms, composed of solid-solution fcc phase, on the surface of a
Fe±Cr±Ni stainless steel by low-energy (1 keV), high ¯ux (1 mA/cm2) N-
implantation near 673 K was reported [260,250,247]. Williamson and coworkers
have shed some light on the mechanisms controlling the thickness of the layers with
high nitrogen content [252]. They achieved nitrided layers of several microns within
about one hour. The rapid growth is enabled by the unusually high di€usion rate of
nitrogen under low-energy high ¯ux conditions [252,260]. The use of such ultrahigh
current densities allows irradiations to very high ¯uences (11019 ions/cm2) [261].

3.5. Plasma immersion ion implantation

Plasma immersion ion implantation (PIII) was initially developed to circumvent


the line-of-sight restrictions of ion beam implantations for material modi®cation
[262]. In PIII, the sample is enveloped by a microwave plasma operated at the elec-
tron-cyclotron resonance mode (2.45 GHz) in the desired low-pressure gas-atmo-
sphere. The sample is connected to a pulsed negative voltage of about 45 kV. Ions in
the plasma are accelerated towards and implanted into the sample homogeneously
from all sides [263,264]. This technique works for many elements (Ar, N, C). The
sample heats by the plasma and is usually also temperature controlled. Thus PIII
can be seen as the combination of plasma nitriding and ion implantation. The
in¯uence of the process parameters on the nitrogen take-up of steels by PIII has
been investigated for example by Blawert and coworkers [265].

4. Laser±material interactions

Laser±material interactions are very complex and only in simple cases, the laser
may be seen as a simple heat source. The numerous facets of laser±material inter-
actions have immediately been the focus of physical research after the ®rst operating
laser was built [28,266±272]. Absorption, heating, melting, evaporation, recoil
pressure, piston e€ect, plasma formation, laser supported absorption waves (LSA),
20 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Marangoni convection, and Kelvin±Helmholtz instabilities are among the aspects


and all these may play an important role for the laser nitriding e€ect.
In this chapter we give a short introduction into some types of lasers, sketch the
laser±solid interaction and estimate the temperature pro®le during the nanosecond
irradiations. Laser supported absorption waves and the predominant material
transport processes will also be discussed. The numerical estimates gained in this
chapter not only serve to classify the phenomena, but will be used for a quantitative
analysis of experimental results in Sections 6±9.

4.1. Laser

It took many years from the publication of the basic theoretical principles of the
laser by Einstein in 1917 [273] until the ®rst radiating laser was built by Maiman in
1960 [274]. With the development of powerful lasers their applications in material
treatments were also rapidly developed [275±280]. Today, an uncountable number of
lasers are used in industrial production processes, especially for the treatment of
metals.
The basic principles of a laser (LASER = Light Ampli®cation by Stimulated
Emission of Radiation) are described in [281±284]. One can divide lasers into gas
lasers (e.g., CO2), solid state lasers (e.g., Nd:YAG) (and liquid state laser) or semi-
conductor lasers. These lasers can radiate either in continuous wave (cw) mode or in
pulsed mode. The ®rst laser was a Ruby laser [274], i.e. a solid state laser. Here, the
inversion is produced by optical pumping. Semiconductor lasers are using a poten-
tial (current) to invert the density of states. A gas discharge inverts the density of
states for the gas lasers. The main properties of laser radiation are the high intensity,
high monochromasy, minimal divergence and high coherence [266]. These are the
reasons for the focussability of laser radiation, which lead to the production of huge
irradiances (up to 1021 W/cm2), enough to evaporate any material or even to start a
nuclear fusion reaction [281]. The spatial energy distribution in the laser beam is
mainly depending on the geometry of the laser resonator, on the mirrors and on the
extraction optics [285,286]. Best focussability is given at low transversal electro-
magnetic radiation modes (TEM00, i.e. Gaussian), whereas for surface treatments a
top hat pro®le is preferred, in order to achieve a homogeneous treatment. The most
important lasers for material treatments are the CO2-laser, the Nd:YAG laser and
the excimer laser. Recently, also diode lasers are attracting great interest [287]. The
excimer laser (excimer = excited dimer) is a pulsed high pressure gas laser, using a
noble gas±halogen mixture [288]. The wavelengths l are in the UV region and
depend on the gas mixture (Ar±F: 193 nm, Kr±F: 248 nm, Xe±Cl: 308 nm, Xe±F:
351 nm [289]). The maximal power is ranging up to 100 MW for pulse durations p
of 10±100 ns and pulse repetition rates of up to fp ˆ 1000 Hz [284,289,290].

4.2. Laser±material interactions

The laser is usually seen as a very special and intense heat source and only its
thermal e€ects are to be considered. Nevertheless, in recent years also laser induced
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 21

chemical reactions, where photons with high enough energy may induce directly
chemical reactions, play a more and more important role [271,28].
In the present case, the photon energy of the l ˆ 308 nm UV radiation is only
Eph ˆ 4:02 eV, which is too small to interact directly with the surrounding nitrogen
gas, which has an ionization energy of Eioniz ˆ 15:6 eV and a dissociation energy of
Ediss ˆ 9:8 eV [291]. Also the intensity of the laser irradiation is too small to induce a
gas breakdown, which needs a threshold irradiance of Ibr ˆ 3  1010 W/cm2 [28,292].
Even in the presence of a metal surface [272,293], which lowers this threshold by two
orders of magnitude, no direct gas breakdown will evolve. Thus the laser irradiation
will hit the iron surface unhindered, without absorption in the nitrogen gas.
Light in the UV region normally interacts only with the electrons of a material,
because ions are too heavy to follow the high frequency ®elds [294]. The optical
properties of metals are determined by the free (valence) electrons, because the inner
electrons only weakly interact with the applied electric ®eld. These free electrons are
accelerated in the electrical ®eld and gain energy. Due to the periodic change of the
®eld vector, the oscillating electrons also re-radiate energy, which causes the high
re¯ectivity of metals. The interactions of laser radiation with matter are becoming
complicated if the pulse duration of the laser is approaching the electron collision
frequency [295]. If the pulse duration is long, compared to the collision frequency
(picoseconds), as in our case, the classical Drude theory [296] can be used, where the
complex dielectric index " ˆ "1 ‡ i"2 is given by:

!2p t2c
" 1 ˆ n2 2 ˆ 1 …17†
1 ‡ ! 2 tc

!2p tc
"2 ˆ 2n ˆ ; …18†
! 1 ‡ !2 t2c

with n and  being the refractive index and the extinction coecient. tc is the mean
time between two collisions. The plasma frequency !p is given by the density of free
electrons N with mass m and the dielectric constant "0 :
s
Ne2
!p ˆ …19†
m"0

The re¯ectivity R and the absorption coecient , or the optical absorption length
dopt ˆ 1= , may be obtained via n and  by
…n 1†2 ‡2
Rˆ …20†
…n ‡ 1†2 ‡2

2! 4
ˆ ˆ …21†
c l
22 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Simultaneously the plasma frequency !p is connected to the speci®c electrical resis-


tivity of the metal el :

1 Ne2 tc
ˆ ˆ !2p tc "0 …22†
el m

For optical wavelengths !  1=tc the re¯ectivity R and the absorption coecient
can be estimated by the speci®c resistance (Hagen±Rubens-equation [294,297]):
p
R ˆ 1 2 2!"0 el …23†
s
1
ˆ …24†
2!"0 el

This explains the high re¯ectivity of metals below the plasma frequency:
R  90 99%, and 1  10 nm [294]. The absorption coecient increases with
decreasing wavelength and is proportional to the speci®c resistance el [9]. Above the
plasma frequency the re¯ectivity drops drastically (UV-transparency of metals)
[298].
The absorption at surfaces not only depends on the wavelength of the laser
radiation, but also on other factors such as incidence angle (Brewster angle), surface
roughness, intensity and temperature of the solid. For example the roughening of
the surface (Ra > l), enhances the absorption by multiple re¯ections [9]. For most
metals the absorption increases with increasing surface temperature. A dramatic
increase of the absorption is found for most metals at the melting point [294]. Also
for higher laser intensities (105±106 W/cm2), anomalous absorption by nonlinear
processes enhances the energy transfer [283,299].
The absorbed energy is distributed by heat conduction. Thus, the heat conduction
equation has to be solved. Since in the present work the dimensions of the laser spot
are much larger than the thermal di€usion length zth , one can use here the one
dimensional equation. For the isotropic case with temperature dependent material
properties this may be written as
 
@T…z; t† @ @T…z; t†
…T†cp …T† l…T† ˆ A…z; t†
@t @z @z
 
@T…z; t† @ @T…z; t† A…z; t†
) k…T† ˆ …25†
@t @z @z …T†cp …T†

with

T(z, t) temperature at depth z at time t


 mass density
k thermal di€usivity, k ˆ clp
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 23

l speci®c heat conductivity


cp speci®c heat capacity
A(z, t) absorbed or released energy per time and volume unitA(z, t) is composed of
the absorbed laser energy Iabs …z; t† and internal heat sinks U…z; t† (phase
transformations).

A…z; t† ˆ Iabs …z; t† ‡ U…z; t† …26†

The absorbed laser energy can be described with the laser intensity I(t), the re¯ec-
tivity R and the absorption coecient :

Iabs ˆ I…t†…1 R† exp… z†: …27†

The temporal shape of the laser beam is assumed to be Gaussian:



I…t† ˆ I0  exp …t t0 †2 =2 2 : …28†
p
In the present case,  ˆ 23:4 ns (55 ns FWHM), t0 ˆ 3, and I0 ˆ H= 2 ˆ 6:8 
107 W/cm2 for H ˆ 4 J/cm2. The material properties generally strongly depend on
the temperature and also internal heat sinks or heat sources have to be taken into
account (phase transformations, melting enthalphy). The absorption depends on the
temperature, but as no data are available, the measured room temperature value is
assumed for all temperatures. The re¯ectivity for iron samples at the laser wave-
length of 308 nm was measured to be 0.44 and the absorption coecient is 1= ˆ 12
nm [291,28]. For 248 nm the values are R ˆ 0:51 and 1= ˆ 16 nm [300,291,28].

4.3. Temperature pro®les

In this section a numerical calculation of the laser induced temperature rise is


performed by the method of ®nite di€erences. Due to the extremely fast and local
supply of the energy the metal can melt and even evaporate. The latter is normally
connected to the formation of a plasma above the surface. The time and laser energy
dependence of the temperature, melt depth and evaporation rate will be calculated.
Since there is no analytical solution of Eq. (25) with temperature dependent
material properties, the calculation has been performed by the method of ®nite dif-
ferences. The problem was treated in one dimension because the laser spot is much
larger than the heat in¯uenced depth. Material transport and convection in the melt
were neglected at this stage. Then, time and space are divided into discrete points
ti ˆ i  t and zn ˆ n  z. Thus, the sample was divided into layers zk (n ˆ 0; . . . ; N)
of distance z and the time steps are t. The partial derivation in time can then be
expressed by the ®nite di€erence (forward di€erentiation) representation:

@T T i‡1 T i
ˆ ; …29†
@t Dt
24 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

where T in is the temperature at the time ti ˆ i  t in the layer zn ˆ n  z. The cal-
culation starts at i ˆ 0 with T 0n ˆ 300 K for all n. The second derivative in space can
also be expressed by a ®nite di€erence

@2 T T n 1 2Tn ‡ Tn‡1
ˆ …30†
@z2 …z†2

Eq. (25) can be re-written as


" #
kin  Ai
T i‡1
n ˆ T in ‡ t   T in 1 2T in ‡ T in‡1 ‡ in † …31†
…z†2 cp; n

which is called the forward time centered space scheme. We have chosen z ˆ 7:5 nm
and t ˆ 0:5 ps, and these values satisfy the Neumann stability criterion …z†2 =t  12 k.
The thermophysical properties of iron needed for the calculations were taken from
Touloukian [301±303], e.g. the thermal conductivity l for iron is shown in Fig. 8 as a
function of temperature.
If the temperature reaches the melting point, signi®cant evaporation starts, and
this must be included into the calculation, because evaporation removes not only
material but also a lot of energy. Above the melting temperature, material is eva-
porated according to the evaporation rate [304,272]:

@zev pD …T†
ˆ p …32†
@t  …2kB T=Ma †

where pD …T† is the vapor pressure of the material at temperature T and Ma is the
atomic mass of the material. Boiling is neglected, since the high recoil pressure and

Fig. 8. Thermal conductivity l of iron (after Ref. [302]).


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 25

the plasma pressure, as we will see later, are increasing the boiling temperature.
Already a plasma pressure of pp ˆ 1:9  107 Pa increases the boiling temperature of
iron to 4730 K [291]. Balandin [305] reported that during irradiation of free iron
®lms of 100 nm thickness with a pulsed Nd:YAG-laser (l ˆ 532 nm,  ˆ 20 ns) no
signi®cant evaporation was observed, although the surface temperature was simu-
lated to reach 4100 K. Another reason is that heterogeneous nucleation is hindered
for heating rates above 109 K/s. Thus boiling is prevented for our pulsed laser irra-
diation [236,306,307]. The iron vapor pressure pD can be described by the equation
of Clausius±Clapeyron
  
Lb 1 1
pD …T† ˆ p0 exp …33†
R T Tb

with the pressure p0 ˆ 1013 hPa, the latent heat of boiling Lb ˆ 349:6 kJ/mol and
the boiling temperature Tb ˆ 3023 K. Then the heat taken away by the evaporation
is given by
@zev
U ˆ    Lb …T†: …34†
@t
The slight dependence of Lb on the temperature is given in [308], where also analy-
tical expressions for the vapor pressure pD and the speci®c heat capacity cp can be
found. The weak temperature dependence of the density  is neglected here and the
value of 7 g/cm3 for liquid iron is taken for the calculation. The latent heat of
melting Lm ˆ 1:62  104 J/mol [291] was used for the calculation. Changes in the
absorption coecient due to the temperature rise were not included [50,75,55].
The absorption of the incoming laser beam is dicult to describe. Song [309] cal-
culated the absorption in the laser produced plasma above a Nickel surface and he
found a transmission of around 90% for 70 MW/cm2. Due to the evaporation of
material from the laser heated surface a `Knudsen' layer is formed above the surface.
These e€ects of laser sputtering, including gas dynamics and re-condensation [310±
312], thermal models [313], and even explosive mechanisms of laser sputtering [314]
are much to complicated to be included in this simple simulation. For simpli®cation,
the absorption in the `Knudsen' [315] is described with the same optical properties as
the Fe bulk material and neglected elsewhere in the plasma, leading to the following
heat source term

Iabs …z; t† ˆ I…t†…1 R† exp… z† exp… zev …t†† …35†

All these details were implemented on a C++ program running on a standard PC


which calculates the temperature pro®le, the molten depth and the evapora-
tion[316,55,50,75]. The program was tested by comparison with results published by
Singh [317] and excellent agreement was found. The results for the present work are
given below. All calculations thereby neglect a possible plasma formation and rela-
ted phenomena, which is one of the main errors of the simulation, besides the con-
stant absorption coecient.
26 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Fig. 9 shows the simulated surface temperature pro®les Ts …t† for the irradiation of
iron with di€erent laser ¯uences H. The bold line represents the pro®le for the
standard ¯uence of 4 J/cm2, for which a maximum heating rate of dT=dt ˆ 100 K/ns
= 1011 K/s and a maximum cooling rate of 50 K/ns = 5  1010 K/s were
deducted. Above 6 J/cm2 the shape of the temperature pro®le is changing, which is
mainly caused by the stronger evaporation from the surface and the sharp drop of
the thermal conductivity (see Fig. 8).
This is also seen in Fig. 10, showing the maximum surface temperature Tmax .
Surface melting starts at about 1 J/cm2, and the boiling temperature Tb is reached
for about 2.4 J/cm2. The increase in temperature with H is slowed down, as the

Fig. 9. Temperature pro®les calculated for laser irradiated iron surfaces. Various laser ¯uences H are shown.

Fig. 10. Maximum surface temperature as calculated for various energy densities H. Vaporization starts
at about Hthr ˆ 2:4 J/cm2.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 27

thermodynamical critical temperature is approached. The corresponding thickness


of the molten surface layer (melting depth) is given in Fig. 11. Again, the thickness
and the duration of the melt increase up to 6 J/cm2, but then saturate at a maximum
depth of about 800 nm, as shown in Fig. 12.
The time during which the surface remains liquid, tmax 2
liq , starts at H = 1 J/cm to
increase linearly with H and reaches its saturation value of about 300 ns for H = 4
J/cm2, i.e. about six times the duration of the laser pulse. Although the boiling
temperature is only reached at 2.4 J/cm2, signi®cant evaporation starts already at
lower laser ¯uences. A layer of 0.1 AÊ is already evaporated at about 2 J/cm2, pro-
viding enough free charge carriers to generate a plasma above the liquid surface. The

Fig. 11. Temporal behavior of the melting depth for various laser ¯uences.

Fig. 12. Maximum melting depth d max


liq for various laser ¯uences H.
28 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

thickness of the evaporated surface layer as function of the laser energy density is
displayed in Fig. 13. About one monolayer is evaporated per pulse for H = 2.4 J/cm2.
Powerful laser beams not only a€ect the intrinsic optical properties but also the
surface topography (shape, roughness) of the irradiated material, which then also
in¯uences the beam±solid coupling. Such surface corrugations are almost always
related to melting or evaporation. Melting of a surface by a laser beam typically
leaves its trace in form of ripples or corrugations. The patterns are often unrelated
to the beam pro®le and appear even if the beam is perfectly smooth [272].
During laser surface treatments the evaporation of material is always combined
with the formation of a plasma above the irradiated surface. Anisimov and cow-
orkers [318] found a nitriding e€ect of irradiated metal and estimated the plasma
pressure via
  !1=5
3… 1† 2=5 30 E2p
pp  ; …36†
4 p6

where is the adiabatic exponent of the gas, 0 is the density of the gas, p is the
pulse duration and Ep is the total energy absorbed by the plasma. Depending on the
laser spot radius rp the high plasma pressure drops according to
r
0
t 0  rp : …37†
2pp

For the present treatment conditions a plasma pressure of about 200 MPa which
lasts for about 260 ns can be estimated with these approximations. Also Mazhukin
[319] discussed temperatures and pressures obtained for irradiation in ambient

Fig. 13. Thickness of the evaporated layer obtained from the simulations with various ¯uences H.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 29

atmosphere. A more detailed treatment of the laser induced plasma above the sur-
face is given in [320], and will be described in the next section.

4.4. Laser supported absorption waves

The vapor formed by intense laser irradiation plays an important role in laser
material treatment. The range of irradiances where evaporation is achieved stretches
from some 103 W/cm2 to the highest realized irradiances of 1021 W/cm2
[269,321,295]. It is clear that many physically distinct regimes are found in this
enormous energy range. At relatively low irradiances (below 106 W/cm2) the vapor is
tenuous and essentially transparent, but with increasing irradiance it becomes
supersaturated. Between roughly 107 and 1010 W/cm2 and depending on the wave-
length, the vapor becomes partially ionized and absorbs a substantial fraction of the
laser energy. On the other hand, radiation re-emitted from the vapor plasma may
heat the solid very eciently [272].
If the vapor becomes ionized and absorbs part or all of the incident irradiation,
the energy is converted into internal energy of the plasma, radiated away as thermal
radiation or consumed in hydrodynamic motion. This plasma forms close to the
evaporating surface and the temperature and degree of ionization depend on the
incident irradiance. If an absorbing gas plasma has formed, an interesting e€ect is
observed. The plasma expands from the surface and moves towards the incoming
laser beam. Such a propagating plasma is called Laser Supported Absorption wave
(LSA wave as visualized in Fig. 14). LSA waves are generally divided into several
regimes: Laser Supported Combustion (LSC), Laser Supported Detonation (LSD),
and Laser Supported Radiation (LSR) (see Figs. 14 and 15). All these and the related
phenomena are extensively described theoretically in the literature [270,272,322±330],

Fig. 14. Principles of a laser supported absorption (LSA) wave. The absorbing region may spread down
to the target surface.
30 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

but experimental results are rarely found. The two most important regimes are
divided according to the propagation velocity of the plasma front, i.e. if the latter is
subsonic or supersonic with respect to the gas. The weakly absorbing subsonic var-
iation is called Laser Supported Combustion wave (LSC). The absorbing plasma
heats and compresses the surrounding gas by expansion and thermal radiation until
this hot and compressed gas itself becomes an absorbing plasma. Under these con-
ditions, the absorption front moves towards the laser beam, because the metal sur-
face blocks its propagation in the opposite direction. For this case a stationary
plasma above the surface is formed. A similar behavior is valid for the LSD wave,
except that there the plasma front is moving with supersonic velocity and the laser
radiation if fully absorbed in the plasma front. The theory of LSC and LSD waves
was formulated by Raizer [270], who calculated the plasma surface pressure (plasma
pressure acting at the surface) caused by the LSD waves to be
  2 1
 2=3 ‡1
pLSD
s ˆ 2 2
1 = ‡ 1 1=3
0 I0
2=3
…38†
2

and the velocity of the LSD wave vLSD


w to be
 1=3
 I0
vLSD
w ˆ 2 2 1 : …39†
0

For the LSC wave [324,235,320] the surface pressure


  1=3  2=3
2W 0 ‡ 1 … 1†… ‡ 1†
pLSC
s ˆ 1 1=3 2=3
0 I0 ; …40†
0 1 2 … ‡ W†… 0 1 2W†

Fig. 15. Plasma pressures acting at the surface: LSC and LSD waves.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 31

with W being a dimensionless particle velocity, W ˆ 0:009  I2=3 2


0 for I0 in MW/cm , is
about ®ve times higher. The wave velocity is
 1=3
2… 1†… 0 1† I0
vLSC
w ˆ … W ‡ 1†  …41†
… 0 ‡ 1† … ‡ W † … 0 1 2W† 

where 0 ˆ 1:4 and ˆ 1:2 are the adiabatic exponents for the surrounding gas and
the metal vapor respectively.
For irradiation with 4 J/cm2, i.e. for an irradiance of 7:27  107 W/cm2, taken as
constant for 55 ns, a LSC pressure of pLSC
s ˆ 48 MPa is derived. This is in agreement
with experimental results given by Schutte [331].
Reilly et al. [332] developed a model for the temporal behaviour of the plasma
pressure in a LSC wave. They estimated the time when the rarefaction fans from the
sides and the top reaches the surface and thus lower the plasma surface pressure.
This is schematically shown in Fig. 16. For the modeling of the temporal behavior of
the plasma pressure acting at the surface, Reilly et al. used a two-dimensional model
[332], which takes into account the expansion of the plasma at its lateral borders.
During expansion, zones with lowered pressures are formed and they move inwards
with the sound velocity cS inside the plasma. The time 2D , which is needed by the
lateral rarefaction fans to reach the center is given by:
rp
2D ˆ ; …42†
cS

Fig. 16. Principles of the decay for a LSC wave: rarefaction fans.
32 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

where rp is the radius of the laser spot (here assumed circular) and the sound velocity
cS is given by
 1=2  1=2
pLSC W ‡ 1 0 1
cLSC
S ˆ s
: …43†
0 W 0 ‡ 1

A second rarefaction wave is starting when the laser pulse ends at time p and the
shock wave is no longer heated by the laser beam, but is still expanding and cooling.
The rarefaction then needs the time z

z ˆ p ‡ vLSC LSC
w p =cS …44†

to reach the metal surface. In conclusion, the development of the plasma pressure
pS …t† in time, acting at the surface, is characterized by the three times p , z and 2D .
If we assume a LSC wave, di€erent cases have to be regarded, depending on the
order of these times. For the present case with the values of I = 72 MW/cm2, 0 ˆ
1:25 g/cm3, ˆ 1:2, and 0 ˆ 1:4 a sound velocity of cS ˆ 7551 m/s is obtained.2
Taking the width of the laser spot as rp ˆ 2 mm, we obtain 2D ˆ 265 ns and from
Eq. (44) follows z ˆ 113 ns. Therefore, the times order as p  z  2D and
according to Reilly [332] the following behaviour of the plasma surface pressure at
the center of the laser spot is obtained:

t  z : p…t† ˆ pP …45†

  2=3
t
z  t  2D : p…t† ˆ pP …46†
z

  6=5
t
2D  t : p…t† ˆ p…2D † …47†
2D

This temporal behaviour is visualized in Fig. 17. The surface pressure in the center
remains constant until z ˆ 113 ns, and then slowly decreases with the time con-
stants z and 2D . It is important to note that the high pressures induced by the
plasma waves is acting for times much longer than the laser pulse duration and also
larger than the time the surface remains liquid.
The light emission of the laser produced plasma is easy to see and the expansion of
the shock wave is easy to hear, so that they can be used to control the laser nitriding
process. No nitriding e€ect has been observed without having the plasma light and
the shock wave detonation.

2
Here, a constant laser irradiance over the 55 ns pulse duration is assumed.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 33

4.5. Material transport phenomena

At laser molten metal surfaces, many mechanisms contribute to material transport


phenomena, such as convection, evaporation and hydrodynamic motions caused by
temperature and pressure gradients [28,272,330,333]. The pressures are produced by
the evaporation itself (recoil pressure) or by the laser supported absorption waves as
discussed before. The most important mechanisms for lateral material transport in
the liquid state are connected to the temperature dependence of the surface tension
 …T† and the piston mechanism [28,268]. Variations of the surface tension may arise
from temperature gradients across the surface of the molten material. If this is due
to an inhomogeneous laser intensity pro®le, this is called thermocapillary e€ect. An
approximation of the radial component of this e€ect is given by BaÈuerle [28]

dliq T d
vlat   ; …48†
v dp dT

where dliq again is the melting depth, T the lateral temperature di€erence, v the
dynamic viscosity of the material, dp the diameter of the laser spot and d=dT the
temperature dependence of the surface tension. An upper limit for the velocity vlat as
calculated with the values d=dT ˆ 5  10 4 N/(m K), dp ˆ 2 mm, dliq ˆ 1 mm,
v ˆ 6:9  10 3 Pa s and T ˆ 4700 K, yields vlat  0:2 m/s.
There are two other main mechanisms of material removal in the beam interaction
zone: (i) melt ejection by the vaporization-induced recoil pressure and plasma pres-
sure and (ii) melt evaporation (high power or short pulses) [334]. At moderate tem-
peratures above the melting temperature the vaporization recoil and plasma
pressure are the primary factors for the material transport out of the laser beam
interaction zone under the regime of hydrodynamic ¯ow. At higher surface

Fig. 17. Temporal development of the plasma pressure (after Ref. [332]).
34 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

temperature (higher ¯uences) the melt removal due to evaporation exceeds the
hydrodynamic mechanism. The mechanisms of the propagation of the evaporation
front were considered in detail by Anisimov and Khokhlov [330] and the results of
the numerical simulations have been given before. An evaporated layer of 8 nm/
pulse has been calculated for 4 J/cm2 (see Fig. 13).
The vapor particles escaping from a hot surface have a Maxwellian velocity dis-
tribution corresponding to the surface temperature, but their velocity vectors all
point away from the surface. This anisotropic distribution is brought to equilibrium
within a few mean-free paths by atomic collisions (Knudsen layer) [315,335±338].
Some of them are also scattered back to the surface and then contribute to the recoil
pressure [334], which is of the order of the saturated vapor pressure [330]. Beyond
this Knudsen layer the vapor reaches a new internal equilibrium with homogeneous
velocity distribution, but with a di€erent temperature. Poprawe [338] made a
detailed calculation of the recoil pressure pr . For irradiation with 4 J/cm2 we
achieved pr ˆ 2:9  107 Pa, when using his parameterizations.
As just discussed, a high plasma pressure and the recoil pressure acting on the
liquid surface inside the laser spot of length a and width b. This pressure di€erence
to the ambient pressure p acts as a piston and moves material from the center
through the sides out of the melt pool. This causes a lowering of the surface by the
piston. This e€ect is shown schematically in Fig. 18.
The problem was treated by von Allmen [268,272] and also by Luft et al. [339] for
laser and pulsed laser drilling. They assumed a non-viscous and incompressible melt
and the pressure distribution was approximated by a `top hat' pro®le with at pres-
sure p0 ‡ p inside the laser spot of radius rp and ambient pressure p0 outside. Then,
the radial velocity of the melt extraction follows from the volume work
s
2p
vlat ˆ …49†


Fig. 18. Schematic view of the piston mechanism (after von Allmen [272]).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 35

where  is the density of the liquid ( ˆ 7 g/cm3) and p the pressure di€erence
which is given by the sum of plasma pressure and recoil pressure p ˆ pp ‡ pr . With
pr ˆ 2:9  107 Pa and pp ˆ 4:8  107 Pa, this yields vlat ˆ 148 m/s, which has to be
compared with a lateral velocity vr  0:2 m/s, induced by the Marangoni convection
[28]. Thus, the piston mechanism should be the dominant mechanism for the lateral
material transport.
The liquid escapes through the circumference of the melt pool and if the two
streams of melt extraction and new laser melting are in a stationary state, i.e. u is
describing the velocity of lowering the piston, we obtain by assuming a rectangular
laser spot with dimensions a  b and the pressure being constant inside the laser
spot:
: :
Vlat ˆ Vpist () 2…a ‡ b†dliq vlat ˆ abu: …50†

The thickness of the melt dliq was estimated [272] to be


 
k Tb
dliq ˆ ln ; …51†
u Tm

so that for the velocity of the piston movement u follows


r  
2…a ‡ b†k ln…Tb =Tm † 2p 1=4
uˆ  : …52†
ab 

Since the numerical simulation for the melting depth dliq has been performed, it is
much more accurate to use this for the calculation of the piston e€ect. From Eq. (50)
we can extract the following expression for the total piston movement or surface
lowering zpist during our laser pulse
… tliq s
2  … a ‡ b †  d liq …t † 2p
zpist ˆ  dt: …53†
0 ab 

By using  ˆ 7 g/cm3, area A ˆ a  b ˆ 2  3 mm2 and the pressure di€erence p ˆ


…4:8 ‡ 2:9†  107 Pa, we calculate for H ˆ 4 J/cm2 with the simulation given above a
piston e€ect of zpist ˆ 12…4† nm/pulse for 4 J/cm2.
Also turbulences or bifurcations may play an important role for a fast material
transport. During carburizing of iron by irradiation with a CO2 laser in propane,
carburized layers of about d  10 mm have been found, where the thickness and
homogeneity of these layers could not be explained with di€usion in the liquid state
alone [340]. Also during the nitri®cation of Ti by irradiation with a ns excimer laser
in nitrogen atmosphere, a signi®cant in¯uence of turbulences for the transport of the
nitrogen is expected [341,342]. These turbulences in the liquid surface may evolve
from pressure gradients, produced by local changes in the plasma density or the
temperature [318]. The number of turns of a turbulence during irradiation is
36 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

approximated [318,343] via the lateral material velocity vlat and the pulse duration
p . It follows a traveling distance s ˆ vm  for a surface element. The lateral exten-
sion abifurc of the bifurcation is approximated by the periodicity of the structures at
the surface after the irradiation [340,343]. For the present case and the velocity vlat ˆ
124 m/s and with p ˆ 55 ns a moving distance of s ˆ 12 mm is approximated.

5. Experimental methods

The following section describes the experimental details of sample preparation,


sample treatment and subsequent analyses via ion beam techniques, MoÈssbauer
spectroscopy, nanoindentation hardness measurements and several other methods
(X-ray di€raction, surface pro®ling, electron microscopy, plasma imaging).

5.1. Materials and sample preparation

Pure iron, a stainless steel and ®ve di€erent commercial carbon steels have been
used as substrate material. The pure iron samples are made of ArmcoTM iron [344]
(purity 99.8+%) and the austenitic stainless steel is 1.4401 (X5CrNiMo18.10.3,
AISI 316, Goodfellow [345]). The plain carbon steels have also been used in the
commercially available state. The typical compositions [346] of these materials are
given in Table 4.
From commercially available rods of 25 mm diameter (Armco and plain carbon
steels) slices of about 1.5 mm thickness were cut. From plates of 1.5 mm thickness
(stainless steel) samples with dimensions of 15  15 mm2 were obtained in the same
manner. Subsequently, these samples were mechanically polished by using ®rst SiC
grinding paper (1200, 2400, 4000 mesh) and then diamond paste with a grain size of

Table 4
Typical compositions of the materials used (all in wt%, Fe balance, according to Ref. [346])

Standard No. C Si Mn P S Cr Mo Ni
(steel symbol)

Fe <0.02 <0.08 <0.02 <0.015


Armco Iron
1.1141 0.12±0.18 0.40 0.3±0.6 0.035 0.035
Ck15
1.1191 0.42±0.50 0.40 0.5±0.8 0.035 0.03
Ck45
1.0601 0.57±0.65 0.40 0.6±0.9 0.045 0.045
C60
1.1525 0.75±0.85 0.10±0.25 0.10±0.25 0.020 0.020
C80 (W1)
1.1545 1.00±1.10 0.10±0.25 0.10±0.25 0.020 0.020
C105 (W1)
1.4401 0.07 1.0 2.0 0.045 0.030 16.5±18.5 2.0±2.5 10.5±13.5
X5CrNiMo18.10.3
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 37

1 mm. The samples were cleaned in acetone after polishing. The resulting mean
roughness Ra is about 10(5) nm for all samples (see also Section 5.6, i.e. the surfaces
appear as a metallic mirror.
Special samples have been prepared in an evaporation chamber for marker
experiments (marker layers). At room temperature, Fe and Au ®lms were alternat-
ingly deposited onto Si wafers (0.5 mm Si(100)) by electron-gun and thermal eva-
poration. The base pressure in the chamber was 6  10 5 Pa, rising to about
5  10 4 Pa during deposition. The thickness and the growth rate of the layers were
measured by a quartz oscillator (microbalance). The growth rate was kept at about
0.3 nm/s [347]. This way, dense and homogeneous metal ®lms can be produced [348].
Here we achieved thin Au marker layers of 10±20 nm thickness, buried in Fe at
depths ranging from 20 to 100 nm.

5.2. Laser treatments

For the laser treatments two di€erent excimer laser systems have been used: a
Siemens XP2020 at the Applikations- und Technikzentrum (ATZ) Vilseck, and a
Lambda Physik EMG202 MSC at the Laser Laboratorium GoÈttingen (LLG). The
XP2020 was used for most of the experiments and for simpli®cation the use of this
laser is always assumed unless otherwise stated. The basic physical properties of the
two lasers are summarized in Table 5. As a ®rst approximation, the variation of the
laser pulse intensity in time can be assumed to be Gaussian. The maximum pulse
energy was not always achievable and was dependent on the actual maintenance
status of the system.
For the laser treatments, the samples were mounted into an irradiation chamber.
This chamber was evacuated (by rotation pump and molecular pump) to about 10 4
Pa. The chamber±pump connection was then closed by a valve and the chamber was
®lled with the respective atmosphere at the desired pressure. Standard gas bottles
were used for the ®lling. Natural nitrogen gas 15 N2 (i.e., with the natural 15 N2
abundance of 0.37 at.%) with a purity of 99.999% was the simplest choice. The
enriched nitrogen gas was pre-mixed from pure, isotopically enriched 15 N2 (99.5
at.%) and natural nitrogen (99.999 at.%) and also stored in standard gas bottles.
The mixture ratio was 1:5 so that the resulting 15 N2 content was c15
N ˆ …16:89  0:14†

Table 5
Basic features of the excimer lasers used

Property Siemens XP2020 Lambda EMG202 MSC

Laser gas XeCl KrF


Wavelength l 308 nm 248 nm
Maximum pulse energy Emax 2J 0.3 J
Pulse duration P (FWHM) 55 ns 30 ns
Pulse frequency fP 2 Hz 10 Hz
Raw intensity pro®le, according to producer Rectangular top hat Linear/Gaussian
Raw beam dimension A0 …a0 b0 † 55  45 mm2 26  9 mm2
38 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

at.%, i.e. 45 times the natural concentration. The chamber was constructed to
withstand pressures of 1.5 MPa so that the pressure region of 10 4 Pa  pN  1:5
MPa was accessible for the laser treatments. Nevertheless most of the experiments
were carried out at atmospheric pressure (1013 hPa). A pressure gauge with an
accuracy of about 5 kPa was used to set the desired pressure with the pressure reg-
ulator of the respective gas bottles. For low pressures, the pressure gauges of the
pumping system (Pirani and Penning) were used.
The pulse energy Ep was measured using a pyroelectric joulemeter with a black
ceramic absorption layer (SpectroLas/GenTec with measuring head PEM50K) and
was always determined as a mean of 10 measurements (with standard deviations
usually below 2 mJ). The same optical path (chamber windows, mirrors, lens) was
used for the measurement as for the later irradiation so that all the losses by
absorption were taken into account. The pulse energy was cross checked from time
to time, and the long time stability was usually better than 3%. The laser spot area
AP on the specimen was determined by irradiating photoactive paper and measuring
the blackened exposed spot with a caliper rule. The accuracy of the order of 3%.
The mean energy density H (or laser ¯uence) was then calculated by H ˆ Ep =Ap ,
with an error below 5%. Although the producer claimed a top hat intensity pro®le
of the laser pulse, the exact spatial intensity pro®le of the raw beam of the XP2020
laser was measured with a CCD device and is displayed in Fig. 19. It is seen that the
laser intensity varies signi®cantly over the spot with higher intensities in the center
and lower intensities at the borders. The variations along the short b-axis are more
pronounced than those along the a-axis and resemble a triangular form.
This raw laser beam was focused onto the specimen through a f = 200 mm plane-
concave quartz (fused silica) lens as shown in Fig. 20. The area, and thus the energy
density, was set via the distance specimen's surface Ð lens (lsl ˆ f d) by moving
the lens with a micrometer screw. Almost all irradiations have been carried out with

Fig. 19. Spatial intensity pro®le of the raw laser pulse for the Siemens XP2020 excimer laser (a0 ˆ 55 mm,
b0 ˆ 45 mm), operated at 30 kV.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 39

the specimen being located between the lens and its focal plane (referred to as posi-
tive focus, lsl < f ). Some test experiments have also been made putting the sample
behind the focal plane (negative focus, ll > f ). Thus the laser ¯uence H can be
obtained with the total pulse energy Ep

Ep f 2
Hˆ  ; …54†
a0 b0 d2

where d is the distance of the sample surface from the focal plane f of the lens and a0
and b0 are determined by the aperture (the same for all experiments).
In order to obtain a more homogeneous intensity pro®le, 5  5 crossed cylindrical
lenses (10 mm width and 50 mm length) with a focal length of 50 mm (referred to as
homogenizer) were used. There is exactly one position (45 mm from lens) where the
single spots from the crossed cylindrical lenses superpose to a ¯at and homogeneous
overall top-hat pro®le, as can be seen in Fig. 21. A much more homogeneous
intensity distribution is observed, as compared to the original raw beam of the laser.
The ®xed distance lens-specimen led to a constant area of the laser spot of 5  5
mm2. Here the energy density can only be varied through the pulse energy by
employing an additional variable beam attenuator (semi-transparent mirror, trans-
mission depending on incidence angle).
The whole irradiation chamber was ®xed onto a computer numerical controlled
(CNC) x±y table, allowing programmed movements of the chamber. Irradiations in
air at atmospheric pressure have been performed without the chamber by putting
the specimens directly on the x±y-table. The laser pulses were correlated in time to

Fig. 20. Schematic drawing of the laser beam focusing onto the specimens surface ( f ˆ 200 mm, focal
length of the lens, rectangular aperture a0  b0 ˆ 55  45 mm2, distance between specimen and focal
plane d, changed by moving the lens).
40 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

the movements of the table, i.e. to the specimen. Several irradiation patterns were
applied, which are characterized in Fig. 22. The simplest case (a) is the repeated
irradiation of the same spot on the sample without any movements (Fig. 22(a),
`single spot irradiation'). In order to cover larger areas, scanning movements of the
sample with respect to the laser spot were used. In case (b), the same spot was irra-
diated n times and then the sample was moved by the width of the spot (y ˆ b, and
after ®nishing a row by x ˆ a) so that a full coverage of the surface was achieved
(`tile irradiation'). In case (c) the sample was moved after each pulse by y ˆ b=n,
and after ®nishing a row, the sample was moved by x ˆ a=m, and the movement in
x-direction was reversed, and so on. Such, each area of the sample was treated with
n  m pulses (`meander irradiation'). The energy density H is always given as the

Fig. 21. Spatial laser intensity distribution obtained by using the additional homogenizer (a ˆ b ˆ 5 mm).

Fig. 22. Irradiation patterns used for the laser nitriding experiments: (a) single spot, (b) tile-like (spots
side by side like Dutch tiles) (x ˆ a, y ˆ b), (c) meander-like overlapping, n  m ˆ 4 
3…x ˆ a=n; y ˆ b=m† (the directions and the step-width of the movements are indicated, see also text).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 41

e€ective energy reaching the sample surface, but this is always a mean value which is
calculated by averaging over the whole spot area (H ˆ Ep =Ap ).

5.3. Ion beam analyses

The role of ion beams in materials science is still growing. Besides material treat-
ments, their useful features for the analysis of elemental compositions, elemental
(isotopic) depth pro®les and sometimes also atomic positions are very valuable for
many problems in that ®eld [349,350]. The principles and examples for ion beam
analysis (IBA) are given in [351,352]. The various methods of IBA are based either
on the Coulomb interaction of the ion beam with the target atoms (ions) or on the
induction of nuclear reactions between the target nuclei and the incident energetic
ions [353]. The ®rst is the basis for Rutherford backscattering spectrometry RBS
[354], whereas the second results in nuclear reaction analysis NRA or resonant
nuclear reaction analysis RNRA [355±357,76]. All the ion beam analyses presented
here were performed at the 530 kV heavy ion implanter IONAS in GoÈttingen [358].
There the following IBA methods are available: Rutherford Backscattering Spec-
trometry (RBS), RBS-Channeling, Nuclear Reaction Analysis (NRA), Resonant
Nuclear Reaction Analysis (RNRA), and Particle Induced X-ray Emission (PIXE).
For the present work RBS and RNRA are the most relevant methods; they will be
sketched in greater detail in the following sections.

5.3.1. Rutherford backscattering spectrometry (RBS)


Based on the ®ndings of Rutherford in 1911 [359], after the establishment of
accelerators, detectors and electronics, this method has developed into a standard
analysis technique. Elemental depth pro®ling via the RBS technique is based on the
elastic scattering of monochromatic ions ( particles of 900 keV were used here) at
the Coulomb potential of the target nuclei and on the energy loss of these ions on
their way into and out of the target material. Measuring the energy spectrum of the
backscattered particles allows the determination of elemental depth pro®les on the
nanometer scale. The mass resolution (and thus the elemental/isotope sensitivity)
depends on the kinematic energy transfer from the elastically scattered particles.
Here, Si surface-barrier detectors with an energy resolution of 12±15 keV (FWHM)
[360] are used to record the energy spectra of the backscattered particles.
The principle of RBS is sketched in Fig. 23. For the surface (z = 0) the situation is
quite simple and based on the conservation of momentum and energy. The mea-
sured ®nal energy Ef of an particle with mass Mp ˆ 4:003 elastically backscattered
from a resting target atom with unknown mass Mt , only depends on its initial energy
Ei and the backscattering angle :
h qi2
Mp cos…† ‡ M2t M2p sin2 …†
Ef ˆ k…Mt ; †  Ei ˆ 2  Ei …55†
Mp ‡ Mt

3
All masses are given in atomic mass units, amu.
42 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

where k denotes the kinematic factor for the elastic scattering process and is repre-
sented for the used setting of  ˆ 165 in Fig. 24.
Accordingly, the mass resolution is best in the low mass region, where the curve is
steepest, and becomes poorer for higher masses. Nevertheless, not only the kine-
matic factor is important, but also the yield of backscattered projectiles, i.e. the
probability for backscattering. This is determined by the cross-section of the scat-
tering process. The intensity I (counts per second) of measured a-particles in the
detector with a solid angle dO is given with the area A, the incident intensity I0 and
the atomic density N by:

dI I0 N d
ˆ ; …56†
dO A dO

Fig. 23. Scheme of the RBS method: elastic scattering at a nucleus (left), depth pro®ling (right).

Fig. 24. Kinematic factor k for  ˆ 165 and Mp ˆ 4 (a-particles).


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 43

with the di€erential cross-section d=dO for Rutherford backscattering in the


laboratory system being
q 2
  2
 
2 2 1 Mp =Mt sin…† ‡ cos…†
d Zp Zt e 4
ˆ q 
  2 : …57†
dO 16"0 Ei sin4 …† 1 Mp =Mt sin…†

Eq. (57) can be simpli®ed for Mp  Mt to the well known Rutherford relation
 2
d Z p Z t e2 1
ˆ : …58†
dO 16"0 Ei sin4 …=2†

This means, that the cross-section is low for low Z elements and increases with Z2 ,
which is visualized in Fig. 25. Thus, the advantage of a better mass resolution for
low masses is counterbalanced by the lower cross-sections.
For the depth pro®ling, when the scattering does not occur at the surface but at
depth z, we now have to deal with the energy loss of the projectiles on their way into
and out of the sample as sketched in Fig. 23. The measured particle energy is then

E f ˆ k  …E i Ein † Eout ; …59†

which can be calculated with the (speci®c) energy loss or stopping power dE/dz:
 …d  …d
dE dE dz
Ef ˆ k  Ei …z†dz … z† : …60†
0 dz 0 dz cos…†

Fig. 25. Di€erential cross section d=dO for  ˆ 165 , Ep ˆ 900 keV, and Zp ˆ 2 (a-particles).
44 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

For the calculation of the energy loss, the relative importance of the various inter-
actions between ion and target mostly depend on the ion velocity and the charge
states of ion and target. At ion velocities below the Bohr velocity v0 ˆ e2 =h, i.e. the
velocity of the 1s electron in the hydrogen atom, the ion keeps its electrons or tends
to become neutralized by electron capture. In this regime Ð which is predominant
for RBS and RNRA, the elastic collision, i.e. the nuclear energy loss dominates. Only
at higher ion velocities v  v0 the ion is stripped and the electronic stopping power
dominates. The energy loss has been theoretically treated [361±363] as well as measured
and tabulated for many projectile±target combinations [364] and is often para-
meterized as the stopping cross-section including the atomic density N of the material

1 dE
" … z† ˆ … z† …61†
N dz

Bethe and Bloch [365±369] gave the ®rst expression for the electronic stopping power
     
dE Z2p Z2t e4 2me v2 v2 v2
ˆN ln ln 1 …62†
dz 4"20 v2 me hIi c2 c2

with hIi denoting the ionization potential …hIi  11:5 eV  Zt †, and v the ion
velocity (v  c ! only ®rst term). Nowadays, many semi-empirical models
[364,370,371,372,373] have been developed to improve the accuracy of the para-
meterizations. The electronic stopping power for hydrogen is given by
  "low "high
"e eV=…1015 atoms=cm2 ˆ …63†
"low ‡ "high

"low ˆ C1  EC2 ‡ C3  EC4 …64†


 
C5 C7
"high ˆ ln ‡ C 8 E …65†
EC6 E

where C1 ; . . . ; C8 are ®tting coecients and partly tabulated in [373]. The above
equations are valid in the range of 10 keV/amu to 10 MeV/amu. The nuclear stop-
ping for incident hydrogen is negligible for energies above 10 keV/amu [370]. For
Helium, the electronic stopping power "He e is derived from the stopping power of
protons "He by [364,373]:

"He 2 2 H
e ˆ He  ZHe  "e …66†

where ZHe is the Helium charge and He can be obtained from the simple polynomial
®t
!
X5
2 i
He ˆ 1 exp Di  E : …67†
iˆ0
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 45

The coecients Di are again tabulated in [373]. The nuclear stopping power is cal-
culated with the universal ZBL potential [364] by:
  8:462Zp Zt Mp
"n eV= 1015 at:=cm2 ˆ "0   ; …68†
Mp ‡ Mt Z0:23
p ‡ Z0:23
t

The reduced stopping power "0 is given by

2 ln…E † 3
r
if Er > 30 keV
6 2Er 7
6 7
"0 ˆ 6 7; …69†
4 ln…1 ‡ 1:1383Er † 5
 if Er  30 keV
2 Er ‡ 0:01321E 0:21226
r ‡ 0:19593E0:5
r

where the reduced energy Er is

32:53Mt E
Er ˆ  : …70†
Zp Zt Mp ‡ Mt Z 0:23
p ‡ Z 0:23
t

As many samples contain more than one element, the stopping power often has to
be calculated for a mixture of elements, which is commonly done via Bragg's rule
[374]
X X
"ˆ ci "i with ci ˆ 1 …71†

Bragg's rule assumes that the interactions between ion and target atom are inde-
pendent of the atom's environment. However, the physical and chemical state of the
atom in a compound is observed to have a considerable e€ect on the energy loss.
The deviations from this rule are most pronounced around the maximum of the
stopping power of solid compounds such as oxides, nitrides and hydrocarbons. They
may be as high as 10±20% [373,375]. For compounds such as Fe2O3 or NbN these
deviations are small (<2%) [373,376]. Furthermore, Ziegler's tables also include
corrected values for these nitride and oxide compounds and the improved model of
`cores and bounds' (CAB) allows a better approximation [377,371,364,373]. All these
equations are also included in program packages for the calculation of the transport
of ions in matter (TRIM95 [253] and SRIM [378]).
Since the energy loss is a statistical process there is also an electronic and nuclear
energy loss straggling (Vavilov [379,380], Bohr [381,362,382], Symon [383], Payne
and TschalaÈr [384±386]). One of the ®rst theories was given by Bohr [381,362] (Bohr
straggling).
2
Bohr ˆ 4Z2p Zt e4  N  z: …72†
46 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

p
For low projectile velocities v < v0  Zt , Lindhard and Schar€ re®ned the Bohr
theory to the so called LS-straggling [363]
2 2
LS ˆ Bohr  L…† …73†

with L…† ˆ 0:861=2 0:0083=2 and  ˆ v2 = Zt v20 . Chu [387] as well as Bonderup
and Hvelplund [388] have again re®ned this model via local Fermi velocities and
thus local electron densities, which are not easy to achieve. Fortunately, the LS
straggling deviates from this here only by less than 10% [388].
The depth resolution z is given with the stopping power dE/dx, the detector
resolution det (12±15 keV), and the beam energy spread beam of the incident par-
ticles (beam ˆ 79 eV for 900 keV He2+ [358,389]):
q
dE 2 2 ‡ 2
z ˆ res ˆ beam ‡ LS det …74†
dz
so that the depth resolution in iron is about 12 nm at the surface and decreases to
about 20 nm at a depth of 500 nm [350].
All these equations and parameterizations of density, cross-sections, stopping
powers and energy straggling are included in the various standard software packages
for RBS analyses. There are RUMP [390±392], GISA [393], IBA [394] and RBX
[395] or SIMNRA [396] as examples. The RUMP code has been used throughout
this work to analyze the RBS spectra measured at the IONAS accelerator with 900
keV a-particles at normal incidence to the samples. The beam diameter was about 2
mm, and the current was set in the range of 5±15 nA depending on the target. Three
Si surface barrier detectors (PIPS [360]) with an active area of 25 mm2 at a detection
angle of  ˆ 165 and a distance of about 24 mm to the target, two of them in Cornell
geometry and the remaining in IBM geometry. The solid angle of the detectors was
determined to be O ˆ 3:4 msr. An ampli®er-ADC-multichannel-PC combination
with four ADCs having 1024 channels each, was used for data acquisition and storage.

5.3.2. Resonant nuclear reaction analysis (RNRA)


The unique feature of any technique for nuclear reaction analysis (NRA) is the
fact that it is sensitive to speci®c isotopes. This restricts its general applicability, but
makes it also most powerful and selective for certain cases [397]. Most of the light
stable isotopes (Z < 30) have strong, sharp resonances in the cross-section  …E† for
nuclear reactions induced by light ions at low bombarding energies (E < 3 MeV),
enabling resonant nuclear reaction analysis RNRA.
Nuclear reaction analyses are mostly used when the chemical composition of the
sample is qualitatively known and a particular reaction is used to quantify the
concentration or concentration changes of speci®c elements (isotopes). There are a
few isotopes which have the potential for depth pro®ling: 13 C, 15 N, 18 O, 19 F, 22 Ne,
23
Na, 24 Mg, 26 Mg, 27 Al, 29 Si and 30 Si [398±400].
In principle, NRA is quite similar to RBS, except that as a consequence of nuclear
reactions X(a, b)Y, the incoming and the detected particle do not need to be identical.
Furthermore, the detecting angles and measurement mechanisms are di€erent [401],
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 47

as discussed by Amsel and Lanford [402], Bird [403], Ziegler [404], Deconnick [405],
Mayer and Rimini [382], and Feldmann and Mayer [351]. In the case of particle±
particle reactions, the energy loss and energy straggling of the incident and out-
coming particle have to be regarded as well as the energy dependence of the reac-
tions. The general set-up for NRA is shown in Fig. 26.
The energy of the measured particle b is determined by the energy loss of the
projectile a on its way into the sample, by the reaction parameters, and by the energy
loss of the ejectile on its way out of the sample towards the detector.
 … z=cos in  
Eout …z† ˆ E E0in "in …E…z0 ††N…z0 † dz0 ; Q; Y
0

… z=cos out
"out …E…z0 ††N…z0 † dz0 : …75†
0

Therefore, the depth resolution can sometimes be very poor (straggling in absorber).
But if there are narrow resonances in the reaction cross-section the depth resolution
can be improved by measuring only the reaction products of the resonance reaction.
The theory of RNRA is well described by Maurel [406], Maurel et al. [407], Vick-
ridge [408], and Hirvonen [397]. Nitrogen depth pro®ling with protons is possible
using the reactions 15 N…p; †12 C, 14 N…p; †15 O and 15 N…p; †16 O. The properties of
these reactions at some resonances are summarized in Table 6. The high cross-sec-
tion of the reaction of the 15 N…p; †12 C 0 ˆ 300 mb [398±400] favors its usage for
depth pro®ling.
The resonant reaction 15 N…p; †12 C at about 429 keV was re-assessed by Osipo-
wicz et al. [409] with frozen target measurements, and the resonance energy was
determined with high accuracy to be E0 ˆ 429:57…9† keV with a Breit±Wigner width
of G ˆ 124…17† eV, so that the cross-section has a very narrow resonance:

…G=2†2
 …E† ˆ 0 : …76†
…E E0 †2 ‡…G=2†2

Since the emission of radiation is nearly isotropic [410], the measured g-yield as a
function of the proton energy is given by
…1 …1 …1
Y…Eb † ˆ "det ONp c…z†g…Eb ; E†f…E; E0 ; z† …E0 † dE0 dE dz …77†
0 0 0

where "det is the detection eciency for the detector with the solid angle O and Np is
the number of incident protons.  …E0 † is the resonance cross section as de®ned in Eq.
(76), and c(z) is the isotope concentration at depth z. g…Eb ; E† is the distribution of
the initial beam energy of the IONAS proton beam
 
1 …Eb E†2
g…Ebeam ; E† ˆ p exp 2
…78†
beam 2 2beam
48
P. Schaaf / Progress in Materials Science 47 (2002) 1±161
Fig. 26. General set-up of an NRA measurement (left). The right side shows the case of a (p, ) reaction.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 49

The energy spread beam of the incoming ion energy Ep at the IONAS accelerator is
energy dependent and is given by
4
2
beam …eV† ˆ 2:09  10  Ep 30 keV ‡2:55 …79†

for single charge particles, where Ep is given in (keV) [389]. For ions with higher
charge state, this has to be multiplied by the charge state. For protons at 430 keV a
value of beam ˆ 36 eV results [358,409]. f…E; E0 ; z† is the probability of ®nding a
proton with the incoming energy E at depth z with the actual energy E0 .
„z 0  !
0 2
0 1 E E0 0 "…z † dz
f…E; E ; z† ˆ p exp …80†
z …z† 2 2z …z†2

where the energy spread is given by

z …z†2 ˆ dopp
2 2
‡ stragg …z†: …81†

The energy spread induced by the Doppler e€ect (due to thermal motion of the
target atoms) can also be roughly described by a Gaussian curve with a width of
[411,412]
p
dopp ˆ 4kB Teff E0 Mp =Mt

with Mp =Mt being the mass ratio between beam ion and the isotope investigated
(proton/15 N, m=M ˆ 1=15) and Te€ the e€ective temperature determined by the
Debye model with the Debye temperature D (Fe: D ˆ 470 K, [413]) as given by
… D =T  
1
Teff ˆ 3T…T=D †3 1=2 ‡  3 d; …83†
0 exp… † 1

which for the present case results in dopp ˆ 43 eV (Tmeas ˆ 80 K).


The energy loss is calculated according to the stopping power after the ZBL the-
ory for the proton stopping power [370] (see also Section 5.3.1) and Fig. 27 shows
the energy dependence of the stopping power for protons in iron and nitrogen over a
large energy region. This stopping cross-sections " of protons in Fe and N only

Table 6
Nuclear reactions for nitrogen determination (after Ref. [349])

Reaction Resonance energy, E0 (keV) Half width (FWHM), G (eV) Cross-section, 0 (mb)
15 12
N…p; † C 360 94000 0.03
15
N…p; †12 C 429 124 300
14
N…p; †15 O 278 1600
15
N…p; †16 O 429 124 0.001
50 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

slightly change over the interesting energy range from 425 to 530 keV. There, the
stopping power and the mean values are "Fe ˆ 18:1…7† eV/(1015 at./cm2) and "N ˆ
8:4…4† eV/(1015 at./cm2). For pure iron, this is equivalent with a stopping power of
dE=dz ˆ 153 eV/nm or dz=dE ˆ 6:5 nm/keV. In order to increase the accuracy, the
stopping power was linearized with very good agreement and the values are then (if
the energy is given in keV): "Fe ˆ …27:79…4† E  0:0203…1†† eV/(1015 at./cm2) and
"N ˆ …13:79…3† E  0:0112…1†† eV/(1015 at./cm2).
In order to calculate the energy spread due to the statistical character of the
energy loss, various models have been developed. The straggling may be described
by a formula derived from the transport equation by Vavilov [379], with simpli®ca-
tions given by Maurel [407]. For large depths, the Vavilov distribution approaches
the Gaussian distribution derived by Bohr [362] or Lindhard and Schar€ (Eq. (73))
[363]. Smulders [414] developed a program for the energy loss distribution based on
the program STRAGGL written by Clarckson and Jarmie [415]. In the original
program, a correction is applied for the binding of inner electrons [416], but it was
shown that this correction is obsolete for protons with moderate energies (<1.5
MeV) [417]. For the present work the LS straggling was used and no Lewis e€ect
was observed [418].
The beam energy spread, the resonance width, the Doppler broadening and the
straggling limit the depth resolution of RNRA:

Etot
z ˆ …84†
dE=dz
q
Etot ˆ 2
resonance ‡ beam 2 ‡ dopp2 2
‡ stragg : …85†

Fig. 27. Stopping power for protons in iron and nitrogen as function of their energy.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 51

2
Here, stragg is always calculated using the LS formalism and the stopping power is
assumed to be constant. This behavior is visualized in Fig. 28 for the case of the
15
N…p; †12 C depth pro®ling in iron matrix (with little nitrogen). The depth reso-
lution is better than 1 nm at the surface and changes rapidly with depth to about 20
nm at a depth of about 300 nm. The depth resolution obtained is mainly limited by
the energy straggling of the protons in the material under investigation.
Background and possible interferences have to be subtracted from the measured
yield in order to obtain an accurate nitrogen pro®le [397]. For RNRA, quantitative
results are generally obtained using standard samples containing a well known fraction
of the element to be determined. A detailed discussion of the preparation and use of
reference standards was described by Amsel and Davies [419]. Frequently used
standards are NbN [420], TaN [420] and CrN [50]. Throughout this work, CrN ®lms
prepared by magnetron sputtering have been used as standards [78,178,148]. When
the yield of sample Am Bn is compared to a standard reference with the known isotope
concentration cst , the atomic fraction of the measured element can be obtained from

cst Y"B
cˆ …86†
Yst "st ‡ cst Y…"B "A †

where the depth is calculated with


… E0
1
z…Eb † ˆ dE: …87†
Eb "…E†

Thus the depth and concentration can be determined successively by

2…Ei Ei 1 †
z…Ei † ˆ z…Ei 1 † ‡ ; …88†
"…Ei † ‡ "…Ei 1 †

Fig. 28. Depth resolution of RNRA (here 15 N…p; †12 C in iron at IONAS).
52 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

where the stopping power is calculated with Bragg's rule according to the actual
concentration at point i. The depth can be converted from (at./cm2) to a nanometer
scale if the density is known.
For the present work, the nitrogen depth and lateral pro®les were obtained using
RNRA via the 429.57(9) keV resonance of the reaction 15 N…p; †12 C.
The narrow resonance width of only 124(17) eV [409] allowed the investigation of
nitrogen depth and lateral pro®les with high accuracy due to the well de®ned proton
beam of IONAS [358]. The available maximum proton energy limits the analysis
depth to about 650 nm. During the measurements the samples were cooled to 80 K
in order to avoid changes due to di€usion or annealing e€ects induced by the ana-
lyzing proton beam of 0.5±1 mA. The diameter of the proton beam was set to 1 or 2
mm so that it was smaller than the laser spot dimensions, but unfortunately too
large to resolve ®ner details of the nitrogen distribution across the irradiated spot.
The lateral RNRA scans were taken at a ®xed proton energy of 432.5 keV corre-
sponding to a depth of about 20 nm in pure iron, by moving the sample relative to
the proton beam in steps of 0.1 mm. The g-spectra were taken by means of a 12 cm
long NaI-detector with a diameter of 16 cm. Its eciency "det for the determination
of absolute nitrogen contents was calibrated with stoichiometric CrN reference
samples prepared by magnetron sputtering. Their nitrogen concentration was
determined by RBS with an accuracy better than 2 at.%.
15 12
Fig. 29 shows a typical spectrum  of the 4.43 MeV g-quanta of the N…p; † C
raw
reaction. The yield curve Y Ep was obtained by integration over the range Es to
Ee also including the ®rst and second escape peaks. The measured yield curves were
corrected by several independent
 background measurements B well below reso-
nance: Y Ep ˆ Yraw Ep B . Thus, all other contributions of parasitic reactions
below the utilized reactions are removed.

Fig. 29. Typical spectrum of a RNRA measurement, Ep ˆ 440 keV, Q ˆ 20 mC.


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 53

Such a corrected curve of the integrated yield of the 4.43 MeV g-quanta versus the
incident proton energy Y Ep is displayed in Fig. 30. Several computer programs
[407,414,408,421±426] have been developed to extract the accurate depth pro®le
from these yield curves. Since there are several mathematically correct solutions for
the deconvolution problem, physically reasonable, additional boundary conditions
(positivity, smoothness) were used here. The measured yield curves were trans-
formed into nitrogen depth pro®les cN …z† by the simulation and ®t program
WinRNRA [427].
WinRNRA deconvolutes the measured yield curve by an iterative method using
Eq. (77), the stopping power as given by [364] including the Bragg rule, the LS-
Straggling, the IONAS beam spread, the Doppler broadening and the atomic den-
sities of the Fe±N system as given in Chapter 2. The latter is also used for the con-
version ofPthe area density tz to a nanometer depth scale z ˆ tz =N. The minimization
of 2 ˆ …Ysim …Ei † Ymeas …Ei ††2 is done by the Levenberg-Marquardt formalism
[428]. The nitrogen concentrations are calculated assuming the natural abundance of
15
N (0.37 at.%) or the actually used enrichment. The calibration sample contained
the natural abundance. If the concentrations are given as 15 N concentration, the
calculation is performed assuming an atmosphere of pure 15 N (100 at.%).
The measurement shown in Fig. 30 was deconvoluted using the WinRNRA pro-
gram with automatic setting of the layer structure. The obtained nitrogen depth
pro®le is shown in Fig. 31. The program chooses more points at the surface and less
points in deeper layers, according to the increase of the straggling. A smooth depth
pro®le is obtained with the simulation. The solid line shown in the yield curve in
Fig. 30 is calculated using this nitrogen concentration pro®le and an excellent
agreement is observed.


Fig. 30. Normalized yield Y Ep for laser nitrided iron as obtained by a RNRA measurement. The solid
line results from the deconvolution (see text).
54 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

5.4. MoÈssbauer spectroscopy

The recoilless nuclear resonance absorption was discovered by Rudolf MoÈbbauer


in 1957 [429±431]. The principles and applications of the MoÈssbauer e€ect are
described in many textbooks [432±439]. Here only the most important facts and the
experimental setup for this work are described.
If a nucleus in an excited state with mean lifetime N (141 ns for 57 Fe) emits g-
radiation, the energy distribution follows the Breit±Wigner law (Lorentzian):
…G=2†2
I…E† ˆ I0  …89†
…E E0 †2 ‡…G=2†2
where E0 is the transition energy and G ˆ h=
 N is the resonance width (E0 ˆ 14:401 keV,
G ˆ 4:7  10 9 eV for 57 Fe). The momentum of the quantum is hk.  A free atom
emitting a g-quantum will experience a recoil with the energy Er ˆ p2 =2M …Er ˆ
2  10 3 eV for 57 Fe), which is generally much larger then the natural line width G.
If the nucleus is bound in a solid, the whole solid takes the recoil and while then the
mass is about 1020 times larger as compared to an isolated nucleus, the recoil energy
becomes negligible. This enables resonant nuclear emission and absorption, i.e. the
MoÈssbauer e€ect. Nevertheless, the emission (or absorption) may also involve
energy transfer to lattice vibrations (phonons) in the crystal and the chance that the
emission takes place without inducing any phonons is called, in analogy to X-ray
di€raction, Debye±Waller factor fD . This factor depends on the temperature T, the
energy of the g-quantum, and the Debye temperature YD of the crystal [435±439]:
"  2 … YD =T #!
3h2 k2 T y
fD …T† ˆ exp 1‡4 dy : …90†
4MkB YD YD 0 exp…y† 1


Fig. 31. Nitrogen depth pro®le cN …z† as obtained by deconvoluting the yield curve Y Ep via WinRNRA.
(The points are indicating the layer structure chosen by the program).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 55

The probability of recoilless emission or absorption is 0.76 for 57 Fe in a-Fe at room


temperature and thus allows fast measurements. Due to this high resolution of
MoÈssbauer spectroscopy, the hyper®ne interactions can be resolved by a Doppler
modulation of the energy
 v
E…v† ˆ E0  1 ‡ …91†
c

by moving the source relative to the absorber. Usually, the source is moved and the
necessary velocities for 57 Fe lie in the range of 10 mm/s.
The high energy resolution of this e€ect allows the observation of the hyper®ne
interactions, i.e. the change of the energy levels due to electric or magnetic ®elds
acting at the nucleus. These are the electric monopole interaction, the electric
quadrupole interaction and the magnetic dipole interaction.
The electrical monopole interaction is the interaction of the nuclear charge Ze
with the electron density at the nucleus [440]. This leads to a change of the energy
states in source (S) and absorber (A) as compared to a point charge [432]:

2 h i
ES ˆ E0 ‡ Ze2 jCS …0†j2 R2e R2g …92†
5

2 h i
EA ˆ E0 ‡ Ze2 jCA …0†j2 R2e R2g …93†
5

Here R is the nucleus radius of the appropriate state (g Ð ground state, e Ð excited
state) and C the electron wave function, ejC…0†j2 then gives the electron density at
the nucleus. In the MoÈssbauer spectra this leads to the isomer shift :

2  h i
 ˆ Ze2 jCA …0†j2 jCS …0†j2 R2e R2g …94†
5

An electrical quadrupole moment of the nucleus, eQ, interacts with the electric ®eld
gradient tensor (EFG) acting at the nucleus, leading to a splitting of the energy
2
levels [440]. Diagonalizing the EFG tensor Vij ˆ @r@i @rj V (V = electrical potential)
with jVzz j  jVyy j  jVxx j and Vxx ‡ Vyy ‡ Vzz ˆ 0, the asymmetry parameter  is
given by:

Vxx Vyy
ˆ …95†
Vzz

and the energy shift for a state with spin I and magnetic quantum number mI is:
r
eQVzz  2  1
E q …m I † ˆ 3mI I…I ‡ 1† 1 ‡ 2 …96†
4I…2I 1† 3
56 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

For 57 Fe with Ig ˆ 1=2 and Ie ˆ 3=2 only the excited state splits and the quadrupole
splitting " is: r
1 1
" ˆ Eq ˆ Eq …mI ˆ 3=2† Eq …mI ˆ 1=2† ˆ eQVzz 1 ‡ 2 …97†
2 3

The magnetic moment  of the nucleus interacts with a magnetic ®eld B acting at
the nucleus (Nuclear Zeeman E€ect). This interaction gives rise to the splitting of
the degenerated states. The state with spin I splits into 2I ‡ 1 magnetic substates
with the eigenvalues [440]

BmI
Em ˆ ˆ gN N BmI …98†
I

where the magnetic quantum number mI can have the values mI ˆ I,


I 1; . . . ; …I 1†; I. The magnetic moment  is given by the Bohr magneton N
and the Lande factor gN by:

 ˆ gN N I: …99†

Based on the selection rule m ˆ 0, 1 only six transitions are possible, leading to
the signi®cant sextets in the MoÈssbauer spectra of magnetic materials.
If the magnetic dipole interaction and electrical quadrupole interaction act toge-
ther [441], the eigenvalues cannot be determined as easily as for each interaction
alone. Normally perturbation methods are applied. As usually the magnetic inter-
action is much stronger, ®rst the eigenvalues are determined for this magnetic
interaction and the electrical interaction is treated as a perturbation. For the energy
states in a ®rst order perturbation calculation one ®nds [442]:
1 1 1 
Em ˆ gN N HmI ‡ … 1†jmI j‡ 2   eQVzz  3 cos2  1 ‡  sin2  cos 2
4 2
…100†
where  and  are the angles between the direction of the magnetic ®eld and the
principal axis of the electric ®eld gradient component Vzz . For 57 Fe we have:
1 
" ˆ eQVzz 3 cos2  1 ‡  sin2  cos 2 : …101†
8
Now the line positions for the general case can be given by Eq. (102).

L1 ˆ  g 1 N H ‡ "
L2 ˆ  g 2 N H "
L3 ˆ  g 3 N H "
L 4 ˆ  ‡ g 3 N H "
L 5 ˆ  ‡ g 2 N H "
L6 ˆ  ‡ g1 N H ‡ ": …102†
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 57

The g-factors gi result from the LandeÂ-factors of the ground state gg and the excited
state ge :
1 
g1 ˆ  3jge j ‡ gg
2
1 
g2 ˆ  jge j ‡ gg
2
1 
g3 ˆ  jge j gg …103†
2
 
The relative line intensities Ii for a sextet depend on the angle  ˆ € k~; I~ between
the propagation direction k~ of the g-radiation and the direction of the spin (mag-
netization) I~ and are given in Table 7.
For  ˆ 90 the intensity ratio is I1 :I2 :I3 ˆ 3: 4:1, whereas for  ˆ 0 the lines 2
and 5 disappear and the ratio is I1 :I2 :I3 ˆ 3:0:1. If the directions of magnetization
are randomly (isotropic) distributed ( ˆ 54:7 ), the ratios are I1 :I2 :I3 ˆ 3:2:1. All
these hyper®ne interactions, except the isomer shift, are also found for Perturbed
Angular Correlation (PAC) [443] and they can be easily converted from MoÈssbauer
to PAC and vice versa [444].
The standard geometry of MoÈssbauer measurements is the transmission geometry
(TMS). The sample is irradiated and the absorption is measured with a detector
behind the sample. The e€ective thickness of the absorber (sample) tA is given [440]
by:
tA ˆ 0 fA NA dA aA ; …104†

with the cross section 0 , the recoilless fraction (Debye±Waller-factor) fA , the atomic
density NA , the absorber thickness dA , and the abundance of the MoÈssbauer isotope
aA . For 57 Fe at room temperature we have the following values: 0 ˆ 256:6  10 20
cm2, fA  0:75, NA ˆ 8:476  1022 atoms/cm3, aA ˆ 2:19% [440]. The e€ective
thickness should have values around ta  1 [440].
The e€ect of internal conversion plays an important role in MoÈssbauer spectro-
scopy. As a result of internal conversion, conversion electrons, Auger electrons and
conversion X-rays are also emitted and can be used for the measurements [445,446].
If the electrons originating from the internal conversion process are detected, the

Table 7
Relative line intensities for magnetic dipole-interaction (a sextet)

Line Transition mI Intensity



L1 3=2 ! 1=2 +1 I1 ˆ 3=8 1 ‡ cos2 
2
L2 1=2 ! 1=2 0 I2 ˆ 1=2 1 cos 
L3 ‡ 1=2 ! 1=2 1 I3 ˆ 1=8 1 ‡ cos2 
L4 1=2 ! ‡1=2 +1 I4 ˆ 1=8 1 ‡ cos2 
L5 ‡ 1=2 ! ‡1=2 0 I5 ˆ 1=2 1 cos2 
L6 ‡ 3=2 ! ‡1=2 1 I6 ˆ 3=8 1 ‡ cos2 
58 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

method is called CEMS  Conversion Electron MoÈssbauer Spectroscopy [445].


Because the electron cannot penetrate detector windows the samples have to be
mounted inside an appropriate detection system [445,447]. Depending on the mate-
rial the electron can only escape from the sample surface within 10±400 [448].
Therefore, CEMS is especially suited for the investigation of surfaces and thin ®lms,
up to a depth of some 150 nm [449].
Also the conversion X-rays or re-emitted radiation can be used to measure the
MoÈssbauer e€ect in the samples under investigation. This method is named CXMS
 Conversion X-ray MoÈssbauer Spectroscopy. The X-rays can penetrate about 10±
30 mm of most metals [448] and thus the information depth of CXMS is in this
order. For convenient measurements in suitable times special detectors with large
solid angles have been developed [450±454].
Based on the di€erent information depths of the di€erent modi®cations of MoÈss-
bauer spectroscopy, it is especially interesting to measure these di€erent radiations
simultaneously. Simultaneous measurements enable to achieve a discrete depth
pro®le of the measured samples. Nasu and Gonser [455] have ®rst applied simulta-
neous CEMS and TMS measurements. Schaaf et al. [456±458] have developed an
apparatus (shown in Fig. 32) for the simultaneous measurement of all three

Fig. 32. Experimental setup for STMRS measurements (Ref. [456]).


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 59

modi®cations and this was named STRMS  Simultaneous Triple Radiation


MoÈssbauer Spectroscopy [457]. It should be noted again that the sampling range in
the case of iron-based materials for the CEMS measurements is about 150 nm deep,
while for CXMS it is 10±20 mm [448,449] and TMS scans the whole sample.
The phase analyses presented in this work were carried out by means of CEMS
and CXMS. The CEMS and CXMS spectra were taken at room temperature by a
STRMS set-up [456,457] with a 57 Co=Rh source (400 MBq) and a constant accel-
eration drive. The conversion and Auger electrons were detected in a He/CH4 gas-
¯ow proportional counter [457,445,447,458]; X-rays following conversion were
detected in a toroidal Ar/CH4 gas-¯ow proportional counter [451,452,454]. The
spectra were stored in a multichannel scaler with a resolution of 1024 channels [459].
Calibration was performed using a 25 mm a-Fe foil at room temperature (a-Fe:  ˆ
0 mm/s, Bhf ˆ 33:0 T). All isomer shifts are related to a-Fe. The spectra were ®tted
according to a least squares routine [460,461] by superimposing Lorentzian lines.
As discussed above, a single line has to be ®tted with 3 parameters (I0 , , G),
whereas a doublet has four parameters4 (I0 , , G, ") and a sextet is ®tted with eight
parameters5 (I0 , , G, Bhf , I21 , I31 , G13 ), where the ratio of the linewidth for the inner
and outer part takes into account additional variations due to small ¯uctuations in
the hyper®ne ®eld, which are more e€ective for the outer lines. Nevertheless this
parameter has usually a value close to 1.0 and can be ®xed if necessary. Additional
constraints are usually introduced to reduce the number of free parameters for the
®tting, i.e. normally the intensity and line width ratios in magnetic phases are
assumed to be equal as well as the linewidth for the subspectra of the same phase. If
necessary, known relations of the relative fractions of the subspectra for a phase are
also constraint (e.g., fa :fb :fc ˆ 1:2:1 for a00 -Fe16N2 and g0 -Fe4N, [44]). As an exam-
ple, the phase 0 has only 12 (7 + 3 + 2) instead of 24 free parameters, which may
facilitate the analysis of complicated spectra with many overlapping phases. As a
further simpli®cation, it is normally assumed that all subspectra and all phases have
the same Debye±Waller factor fD . Thus the phase fraction can be identi®ed with the
relative fractions (areas) of the phases in the MoÈssbauer spectrum.

5.5. Microhardness by nanoindentation measurements

The hardness of a material (or a surface) is de®ned as its resistance against the
penetration of another (harder) material [462]. This de®nition is not completely ful-
®lled for the usual hardness testing methods according to Vickers, Brinell and
Rockwell, because the relevant values for the calculation of the hardness were mea-
sured after removal of the test force. For the universal hardness testing, also called
registering hardness testing, these values are determined while the test force is acting.
This method was developed in the past years and recently reached technical applic-
ability because the accuracy of measuring the test force, the indentation depth, and

4
In single crystals or textured materials, the intensities of the two lines may di€er [435], resulting in an
additional parameter.
5
I21 , I31 are the intensity ratios of lines, accounting for magnetic orientation.
60 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

the data registration have been improved [463±470]. An essential advantage of this
method is the fully automatic measurement of the indentation depth, as compared
to the subjective visual estimation of the indentation area. A further advantage is the
force/identation (d(F)) curve itself, which allows the extraction of additional infor-
mation, e.g. Young modulus, and plastic and elastic contributions. There are three
ranges of universal hardness: the macro range for 2 N  F  1000 N, the micro
range for F < 2 N, and the nano range if the indentation depth remains below 200 nm.
A schematic illustration of the universal hardness measuring method is displayed
in Fig. 33. The indentation depth h is continuously measured during the stepwise
increase and afterwards also the release of the force F. The diamond indenter
(Vickers diamond, i.e. a square based pyramid with an opening angle of 136 )
indents into the surface according to the applied force F. The Universal hardness
HU (normally given in (N/mm2) or (GPa)) is then de®ned as the ratio of force F and
the area A, where the Vickers diamond is in contact with the material, which of
course is a function of the depth h:

F F
HU ˆ ˆ : …105†
A…h† 26:43  h2

This universal hardness can be approximately converted into a Vickers hardness


value by dividing the HU (N/mm2) value by 9.81 (HV = HU/9.81, (N) ! (kg)).
Since all real Vickers diamonds have some deviations from the ideal shape of a
pyramid, this may lead to huge errors in the hardness especially for low depths.
Therefore, the area A…h† of the diamond is individually corrected by a correction
measurement of 5 mm thick Si(100) crystal with well known hardness related para-
meters and stored in the computer for the analysis. In order to state precise results,
also the testing force (and sometimes also the testing time and number of steps) are
speci®ed, i.e. if the testing force is 1 N and the hardness was 8500 N/mm2, the proper
citation is: HU 1 = 8500 N/mm2 (=8.5 GPa, 866 HV 0.1).

Fig. 33. Example measurement and scheme for nanoindentation hardness measurement.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 61

Additionally the elastic and plastic work can be determined as well as the Young
modulus E can be obtained from the indentation curves. Using calculations from
Sneddon [471], Loubet [472] has derived a correlation between the Youngs modulus
E and the maximum force Fmax and the diagonal D of the remaining indenter
imprint via:
p
 2Fmax
E= 1 2 ˆ …106†
2D…hmax hR0 †

where  is the Poisson number, hR is the depth after complete force release, and hR0 is
determined by the tangent on the force release curve as a measure of the fully elastic
property given there.
All the values for the microhardness or the laser treated samples given here, were
measured by the nanoindentation method employing a Fischerscope HV100 with a
Vickers diamond [470,464]. The force can be applied with an accuracy of 20 nN and
the depth can be measured with a resolution of better than 2 nm. The maximum
indentation depth is 700 mm. The surface (zero depth) is determined by a special
approaching procedure with an accuracy depending on the surface roughness and
being in the order of nanometers. We utilized the load force F ˆ 4 mN applied in 85
steps with a holding time of 0.1 s for each step for the determination of the surface
hardness. For the layer hardness (surface hardness), the indentation depth should
not exceed 10% of the layer thickness [462]. For the hardness depth pro®les, the
maximum load of 1 N was applied in 150 steps, also with a holding time of 0.1 s for
each step. The maximum indentation depth here is several micrometers, depending
on the hardness of the material.

5.6. Surface pro®ling

Surface investigation of the samples was done by means of optical microscopy and
two pro®lometers (a Veeco DEKTAK 3030 Auto II and a DEKTAK3ST). The
former uses a diamond tip of 12.5 mm radius and a load of 10 mg, the latter o€ers
better resolution by using a 2.5 mm diameter diamond tip. The mean roughness of a
surface with the surface pro®le h…y† is de®ned with the mean height (zero line) h the
measured length l by [473,474]:
…l
1
Ra ˆ jh…y† h j dy: …107†
l 0

5.7. Plasma imaging

With the aim to achieve a deeper insight into the temporal and lateral evolution of
the laser produced plasma, ultra-fast cameras have been used to follow the plasma
behaviour. First experiments in this direction have been carried out by Queitsch and
Schutte [475,331] and Mele et al. [476]. For the present case the plasma photography
62 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

was carried out employing a NANOGATE system (PCO Computer Optics) with an
integrated ampli®er, coupled with a standard CCD camera. The ampli®er consists of
a photocathode, a micro-channel-plate and a phosphorus screen. The spectral sen-
sitive range was between 200 nm and 750 nm with the maximum located at 450 nm.
The CCD camera had a lateral resolution of 512  512 pixel. The shortest possible
exposure time is 10 ns. The camera was triggered by the laser pulse and the time
delay between the start of the laser pulse and the start of the picture collecting could
be varied between 10 ns and 1 ms with an accuracy of about 5 ns. Both the time
delay and the exposure time can be varied in steps of 1 ns. The pictures were taken
facing the short side of the laser spot (b-side) perpendicular to the surface normal
and the laser incidence. The chamber walls were covered with black paint in order to
avoid re¯ections from the metallic walls. Since the dynamic range of the camera was
only 64 grey scales, the ampli®cation and exposure time had to be adapted to the
plasma intensity in order to make all the details of the plasma development visible,
but loosing comparability. In order to provide comparability, the ampli®cation and
exposure time were ®xed, but then it was not possible to resolve all the details of the
plasma.

5.8. Further analyzing methods

For selected samples X-Ray Di€raction (XRD) was applied as a complementary


method for phase analysis. A Philipps PW1740 and a Bruker AXS D8 Advance with
GoÈbel mirror were used for the measurement of the di€raction patterns. Both dif-
fractometers employed Cu-Ka radiation. Measurements with the Bragg±Brentano
geometry and, for better surface signi®cance, also measurements at a ®xed incidence
angle of 5 were performed (i.e. an information depth of 1.5±3.5 mm). Besides stan-
dard XRD data collections [477±480] and standard routines [481], also the Rietveld
program [482±485], was used for data analysis and structure determination.
In addition, metallography, Glow Discharge Optical Spectroscopy (GDOS),
Scanning Electron Microscopy (SEM), Transmission Electron Microscopy (TEM),
and Energy Dispersive X-ray Analysis (EDX) have been employed for the char-
acterization of the nitriding results in selected samples.
A LEO 982 high resolution SEM (HRSEM) was used for image analysis. A SEM
with EDX (Leica 5360, Fa. Leitz) was used for image and elemental analysis via
EDX and also for the analysis of brittle fractures. The brittle fracture samples were
tilted by 15 . A JEOL FX2000 STEM with an acceleration voltage of 200 kV was
used for ®rst TEM investigations.

6. Raw beam irradiations

In this chapter the results of multiple single spot irradiations of pure iron with the
focused raw beam of the excimer laser are described. All treatments were carried out
in pure nitrogen atmosphere of 1013 hPa. By combining the results of the various
methods discussed in Section 5, a detailed scenario concerning the nitrogen pro®les
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 63

and phases, surface structure, material transport and hardness has been developed.
Variations of the laser energy density and intensity pro®les have also been studied.
Fig. 34 shows the top view of the laser irradiated iron surface. The following
nitrogen depth pro®les are measured in the center of the spot at point C and the
lateral nitrogen pro®les, the surface pro®les and the lateral microhardness pro®les
are measured along the line A±B. Further details of the picture are discussed
according to the points indicated with a±d, w, and f.
It is clearly visible that the laser irradiation led to signi®cant changes at the sur-
face. There is the inner rectangular part marked by a black line, which coincides
with the laser spot. The inner part is dark and the surface becomes brighter
approaching the borders of the spot. But also outside the laser spot changes are
visible.

6.1. Surface pro®les

In Fig. 35, taken with better resolution at point w, laser induced surface corruga-
tions are visible with a periodicity of about 30 mm. A peak to valley height of about
2 mm is observed [50,55]. The `waves' seem to be arranged concentric around the
center, where the corrugations are less pronounced. They also become weaker
towards the border of the spot. These surface structures with a periodicity con-
siderably exceeding the wavelength of the laser frequently occur if the energy density

Fig. 34. Optical image (topview) of iron irradiated with 64 pulses of the raw beam hitting the same spot
(fp ˆ 2 Hz).
64 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

is high enough to produce a plasma over the molten surface [486±488]. Ang [489]
found similar wave structures in the order of 30 mm on aluminum. Recently these
structures were postulated to be caused by the Kelvin±Helmholtz instability [490] at
the interface between molten layer and plasma plume [489]. The instability is driven
by the radial expansion of the plasma plume acting much like wind over water. The
pronounced periodicity of the surface pattern in the case of irradiation with the raw
beam could be a result of the more pronounced radial movement of the plasma
plume. In [489] it is derived that the instability growth rate depends on the plasma
wind kinetic energy. There it was assumed that the most probable period is always
equal to the observed period of 30 mm and that the thickness of the molten layer
does not change considerably with the homogeneity of the laser beam.
Outside the laser spot, at position f, the surface covered with some `fall-out',
which can easily be wiped o€, leading to the observed `scratches'. The fall-out region
is separated from the the laser spot by a white clean area, which can be explained by
geometrical e€ects of the plume expansion [491]. RNRA and MoÈssbauer investiga-
tions proved that this fall-out consists of iron-nitrides (e-nitride) with a nitrogen
concentration of about 30 at.% [50]. This fall-out is shown also in Fig. 36, and the
surface covered with this fall-out resembles the surface of gas-nitrided samples [217]
and it forms some small spherical particles with sizes up to about 100 nm. The fall-
out has only loose contact to the surface. Sometimes particles are partly molten into
the iron surface [75].
The e€ects of the laser irradiation on the surface topology are visible in the surface
pro®le shown in Fig. 37 taken by Illgner et al. [55] with the surface pro®ler along the
line A±B (see Fig. 34).
This pro®le exhibits drastic changes in the surface of the irradiated sample as
compared to the ¯at `zero' line of the untreated material. A `deep' trough of about 5
mm depth in the center of the spot and `steep' ridges of about 6 mm height can be

Fig. 35. Optical image of the surface corrugations evolving after irradiation with 64 pulses at 4 J/cm2,
taken at point w.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 65

seen. The ridges lie well inside the laser spot. If we have a closer look at the outer
part of the pro®le, as displayed in Fig. 38, also some changes outside the laser spot
can be observed.
The thickness of the fall-out d fo outside the laser spot decreases exponentially with
the distance x to the spot border according to the equation
 
fo fo x
d …x† ˆ d 0  exp …108†
xd
with the parameters being d0 ˆ 1:3…2† mm and xd ˆ 0:27…2† mm.
This behavior is also re¯ected in the pictures of a brittle fracture taken in the SEM
as shown in Fig. 39. In the picture (a) taken in the center of the spot, the maximum

Fig. 36. SEM picture of the surface at point f showing the fall-out outside the laser spot.

Fig. 37. Pro®lometer scan of iron irradiated with 64 pulses of the raw beam at 4 J/cm2 and 1013 hPa. (The
bar indicates the laser spot dimension.).
66 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Fig. 38. Pro®lometer scan of the border region (point d) with underlying picture. The measurement was
carried out along the lower abscissa.

Fig. 39. Brittle fracture as observed in the SEM, pictures (a)±(d) are taken at the positions a±d in Fig. 34.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 67

melting depth of almost 1 mm nicely ®ts to the calculated depth of dliq ˆ 800 nm
presented in Chapter 4. The layer thickness is increasing at points b and c and the
picture (d) shows the ridge of the crater.
Figs. 37 and 39 clearly demonstrate a material transport from the center towards
the border of the laser spot. This is resembling the piston-e€ect already discussed in
Chapter 4. As a quantitative measure for this material transport, the corresponding
thickness of move material d‡ and d , normalized to the lateral spot dimension
b, were calculated from the surface pro®les h(x) according to
…
1 ‡1 1
d  ˆ  …h…x†  jh…x†j† dx: …109†
b 12

The results are summarized in Fig. 40 for various numbers of pulses n. There is a
linear dependence of d ‡ and d with n, d ‡ ˆ 10…1† nm/pulse) and d ˆ
23…2† nm/pulse. Thus the loss is much higher than the ablated layer thickness of 8
nm/pulse obtained from the simulation. The mean lowering of the surface in the
center of the spot ( 1 mm  x  + 1 mm) is z ˆ 55…14† nm/pulse [48], which is
even more than the theoretically predicted value of 12(4) nm (see Section 4.5).
This lowering of the surface is additionally con®rmed by the cross section TEM
picture shown in Fig. 41. The lateral movement of material from the center leads to
the deeper melting during the next pulse, so that a periodic structure is observable.
The observed periodicity is changing between 50 and 100 nm, in accordance with the
value given before. The losses in the center are higher than the gain at ridges around.
For an accurate balance, however, this has to be measured in two dimensions over
the whole surface including also the fall-out region outside the laser spot [52,55].
Normalizing this material transport to the appropriate surface areas, the loss is
d ˆ 33…4† nm/pulse, whereas the gain in the ridges is d ‡ ˆ 31…4† nm/pulse. Thus

Fig. 40. Material transport d ‡ and d for increasing number of pulses n.


68 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

it was found, that the total loss of material from the sample was only d ˆ d ‡
d ˆ 2…1† nm/pulse6 [55] close to zero. The thickness of the fall-out was esti-
mated to be 2 nm/pulse. All these results are in good agreement with the simulated
transport due to piston e€ect and evaporation [55,50].
The surface roughness of the laser nitrided surfaces increases with the number of
pulses according to the graph shown in Fig. 42, starting from Ra ˆ 10…5† nm for the

Fig. 41. TEM cross section of a raw beam irradiated sample (64 pulses, 4 J/cm2, stainless steel).

Fig. 42. Surface roughness Ra for the laser nitrided samples (4 J/cm2).

6
The di€erence was determined independently for each sample, therefore, the error is smaller as com-
pared to the mean d ‡; values.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 69

untreated sample to about 2.5(1) mm after treatment


p with 512 pulses. When ®tting
the data points with the function Ra ˆ R0a ‡ R0a  n, the value R0a ˆ 110…10† nm is
obtained.

6.2. Lateral nitrogen pro®les

Lateral nitrogen concentration pro®les were taken by moving the analyzing pro-
ton beam across the laser spot along the line A±B in Fig. 34 at a ®xed proton energy
of Ep ˆ 432:5 keV (20 nm depth). The results are shown in Fig. 43 for various
numbers of laser pulses n at 4 J/cm2.
For the irradiation with a single pulse n ˆ 1 a nearly Gaussian nitrogen pro®le
across the laser spot is observed. The maximum concentration is located in the cen-
ter and amounts to about 11 at.%. When increasing n the maximum concentration
in the center is hardly changed, but close to the spot borders new maxima start to
build up. For n ˆ 64 pulses these maxima come close to 22 at.% and are located just
outside the laser spot. Remembering the results of the previous section, it is imme-
diately clear that this behaviour is related to the deposition of the nitride fall-out just
outside the spot.
The simulation of the lateral RNRA pro®le was based on the lateral thickness
pro®le of the fall-out shown in Fig. 38, the nitrogen concentration of 30 at.% in this
deposited fall-out and a circular proton beam of 2 mm diameter [55]. In addition,
the nitrogen distribution inside the laser-irradiated area was assumed to be equal to
the distribution pro®le determined for the sample irradiated with only a single laser
pulse. This latter assumption is justi®ed by the fact that the nitrogen concentration
in the laser-irradiated area did not change with the second laser pulse. The resulting

Fig. 43. Lateral nitrogen pro®les across the laser spot for various numbers of pulses n at 4 J/cm2 as
obtained by RNRA. Ep ˆ 432:5 keV (20 nm) and rp ˆ 2 mm. The laser spot dimension is marked by the
bar.
70 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

simulated concentration pro®le shown in Fig. 44 reproduces well the measured


pro®le. This result is of particular interest in two respects: scanning the laser beam
across the sample during irradiation would move the fall-out of each pulse into the
center of the irradiated area of the subsequent laser pulse and therefore enhance the
overall nitriding e€ect. Secondly, we note that scanning of the laser beam over the
sample could leave non-adherent fall-out with a high nitrogen concentration on top
of that material which had been irradiated previously. This fall-out could a€ect all
those nitrogen concentration and phase analyses which integrate over larger areas,
including regions outside the laser-irradiated area where the fall-out was detected.
A ®nal comparison of the lateral nitrogen pro®le with the surface pro®le and the
lateral hardness pro®le will be given in the summary.

6.3. Nitrogen depth pro®les

Fig. 45 shows the nitrogen depth pro®les for laser nitrided iron measured in the
center of laser spot as indicated by `C' in Fig. 34. The treatment parameters are again
4 J/cm2 and 1013 hPa nitrogen pressure for n ˆ 1 to 64 pulses. It can be recognized
that all curves show about the same maximum surface concentration of about 30
at.%, independent of the pulse number. At larger depths the nitrogen concentration
is increasing with the pulse number and seems to saturate at about 32±64 pulses.
The nitrogen depth pro®les, as revealed by RNRA, were ®tted with two super-
imposed di€usion pro®les with an o€set according to

cN …z† ˆ c0 ‡ c1  erfc …z=z1 † ‡ …c2 c0 †  erfc…z=z2 † …110†

where z is the depth, c0 the o€set, c1; 2 are the maximum concentrations and z1; 2 the
di€usion lengths. This is an extension of the Eq. (10). The resulting parameters of
the least squares ®ts are summarized in Fig. 46. The parameters c1 , z1 , c2 , z2 turned

Fig. 44. Simulation of the lateral nitrogen pro®le for n ˆ 64 (see text).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 71

Fig. 45. Nitrogen depth pro®les for single spot laser nitriding of iron with increasing number of pulses n
(4 J/cm2, 1013 hPa). The solid lines are ®ts according to Eq. (110).

Fig. 46. Results of the least squares ®ts according to Eq. (110) as shown in Fig. 45.
72 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

out to be independent of n, although parameter c1 shows larger ¯uctuations. Their


mean values are: c1 ˆ 16…2† at.%, z1 ˆ 16…2† nm, c2 ˆ 12…1† at.%, z2 ˆ 109…3† nm.
The parameter c0 saturates at 3.4(2) at.%.
Now these two di€usion pro®les have to be interpreted. A mean temperature of
T ˆ 2445 K over a one-pulse cycle during which the surface is molten (tm ˆ 293 ns)
to a maximum melting depth of dliq ˆ 800 nm can be estimated, based on the tem-
perature calculations presented before. This temperature corresponds to a mean
di€usion constant of D ˆ 1:42  10 4 cm2/s when using the approximationpfor the

di€usion of N in liquid iron as given in Section 2.5, [164]. Thus, via z1; 2 ˆ 4Dt1; 2
the two di€usion times t1 ˆ 4…1† ns and t2 ˆ 209…12† ns are obtained.
The ®rst di€usion pro®le appears too short to be attributed to a real di€usion
process. In fact, it should be attributed to the iron nitride fall-out [55]. Integration
leads to a layer thickness of about 5(1) nm at a mean nitrogen concentration of 30
at.% in the fall-out. The second time constant t2 ˆ 209 ns is well below the 293 ns
for which the surface remains liquid. According to the model of laser supported
combustion (LSC) waves, a high plasma pressure of about 48 MPa above the liquid
surface is generated [270,332,320]. It remains constant for z ˆ 113 ns when the
rarefaction fan from the top reaches the surface, and then slowly drops until 2D ˆ
265 ns when the rarefaction fan from the lateral borders of the laser spot reaches the
center [332]. The di€usion time t2 ˆ 209 ns lies in between these two values. The
nitrogen solubility in the liquid iron surface at the given pressure amounts to about
5 at.% (at T ˆ 2445 K) [492], less than half of the experimental value c2 ˆ 12 at.%.
Nevertheless, the second di€usion pro®le is attributed to an inward di€usion of
nitrogen dissolved in the upmost liquid surface layer, as the plasma-enhanced nitrogen
activity may lead to a solubility that can be three times higher than for molecular N2
[156]. In order to simplify an analytical simulation, the depth pro®les were also ®tted
with two superimposed exponential functions with an o€set according to
   
z  z
cn …z† ˆ An0 ‡ A1 exp ‡ A 2 A n
0 exp …111†
z01 z02
and the corresponding results are summarized in Table 8. The resulting weighted
mean parameters for all n are A1 ˆ 17…2† at.%, z01 ˆ 6…1† nm, A2 ˆ 14…2† at.%, and
z02 ˆ 52…4† nm.

Table 8
Results of the exponential ®ts to the nitrogen depth pro®les. The last line gives the weighted mean values

n A0 (at.%) A1 (at.%) z01 (nm) A2 (at.%) z02 (nm)

1 0.17  0.10 16.6  0.6 4.9  0.4 15.0  0.4 46.9  1.9
2 0.34  0.14 17.5  0.7 6.3  0.5 15.5  0.6 52.7  2.9
4 1.13  0.16 15.8  0.7 8.5  0.9 12.7  0.7 62.4  5.1
16 2.53  0.67 22.0  2.2 8.8  1.8 9.1  2.3 69.7  28.6
32 2.93  0.22 28.6  1.5 5.5  0.5 10.1  0.8 52.6  7.1
64 3.45  0.15 16.0  1.3 7.9  2.2 13.1  1.6 48.8  6.1
  17.1  1.2 6.3 0.9 14.0  1.1 52.5  3.6
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 73

The saturation observed in the nitrogen depth pro®les is re¯ected also in the total
amount of nitrogen found in the ®rst 300 nm. This quantity is given as the mean
nitrogen concentration hcN i, de®ned by

… 300 nm
1
hcN i ˆ  cN …z† dz: …112†
300 nm 0

Also the quantity hcN i shows a steep increase during the ®rst pulses, followed by a
saturation for larger n. The values cmax
N ˆ 5:0…2† at.% and n0 ˆ 1:5…2† are obtained
when ®tting with the activation function hcN i ˆ cmax
N  …1 exp… n=n0 ††. The satura-
tion behavior will be discussed in detail after presenting two more important facts.
One is the material transport and intermixing during the molten state of the surface
and the other is the development of the nitrogen take-up with increasing number of
pulses.

6.4. Irradiation of marker layers

The material transport and intermixing in the melt during the pulse was investi-
gated by the treatment of an iron sample containing a 10 nm thick Au marker layer
at a depth of about 60 nm. This sample, produced by electron-gun evaporation, was
measured with RBS and the spectra are displayed in Fig. 47 together with the Au
concentration pro®les derived form these RBS spectra using the pro®le command of
RUMP rump [390].

Fig. 47. RBS measurement of a Au marker layer: (a) RBS spectra before and after laser irradiation
(4 J/cm2, 1013 hPa N2), (b) Au depth pro®les derived from (a).
74 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

The thin marker layer is visible before the irradiation and can be ®tted with a
narrow Gaussian pro®le. After the laser treatment, there is only a constant Au
concentration down to the visible depth of about 400 nm. So even a single laser
pulse is enough to homogenize the sharp marker layer causing a complete intermix-
ing of the Au with the iron. The Au concentration is further decreasing with the
following pulse as it should be expected due to the lateral transport of the Au con-
taining melt out of the center by the piston mechanism. Unfortunately, due to the
limitation in energy and the overlap of the Au and the Fe signal in the RBS spectra,
the total amount of Au, and thus the loss of Au cannot be determined. This homo-
genization of the Au layer could not be caused by di€usion alone but must partly be
due to convective transport of material. Pure di€usion would cause a broadening of
the Au layer from  ˆ 9…1† nm to  ˆ 62 nm for a single laser shot, when taking the
calculated temperature pro®le and integrate over the time when the surface is liquid
(300 ns). The Au di€usion constant is D ˆ D0 exp… Q=RT† with D0 ˆ 31 cm2/s and
Q ˆ 14:9 kJ/mol [291]. Thus it can be assumed that also the nitrogen dissolved in the
melt is e€ectively transported and homogenized by piston e€ect and convective ¯ow
in addition to the di€usion.

6.5. Isotopic experiments and modeling of depth pro®les

RNRA via the reaction 15 N…p; †12 C is only sensitive to the isotope 15 N. Varying
the 15 N content in the irradiation atmosphere enables additional experiments, which
highlight e€ects of single laser pulses in a series of irradiations. For example, the
nitrogen take-up can be investigated by performing alternating laser treatments in
isotopically enriched atmospheres [54]. In Fig. 48, the 15 N concentration for the
irradiation of iron with only one laser pulse in 15 N enriched atmosphere (16.89 at.%

Fig. 48. 15 N depth pro®les as measured with RNRA with RNRA: (a) one pulse in enriched atmosphere
(0/1), (b) 63 pulses in natural nitrogen followed by one pulse in enriched nitrogen (63/1).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 75

15
N, sample (0/1))7 is compared with the 15 N pro®le after irradiation with 63 pulses
in natural nitrogen (0.37 at.% 15 N) followed by one pulse in the 15 N-enriched
atmosphere (sample (63/1)). With RNRA only the new nitrogen take-up with the
64th pulse is visible for sample 63/1. The solid line is drawn according to Eq. (111)
with the mean values given above. The coincidence of the theoretical line and the
measurements again shows the validity of the approximation. Comparing the con-
centration pro®les of these two samples we see that the take-up of new nitrogen does
not change with the number of pulses, i.e. the take-up of new nitrogen is not changed
with the number of pulses and the saturation behavior should have other origins.
For the simulation of the depth pro®les and their development with n, we now
have to take into account the e€ects shown above and assume the following process
mechanisms in the given order:

1. Loss of a surface layer z = 55 nm/pulse including its nitrogen by ablation


and `piston' e€ect;
2. Complete intermixing of the remaining nitrogen, resulting in a constant nitro-
gen concentration down to the liquid±solid interface at a depth of dliq ˆ 800
nm (no nitrogen below dliq );
3. Constant take-up of new nitrogen, independent of n via inward di€usion from
the surface and fall-out, according to Eq. (111).

Thus the nitrogen concentration after n pulses can be expressed by Eq. (111) where
the o€set for the ®rst pulse should vanish (A10 ˆ 0) and the o€set after (n ‡ 1) pulses
can be generated successively by
… dliq
1
An‡1
0 ˆ cn …z† dz …113†
dliq z

… dliq   
1  
An‡1
0 ˆ  A1 exp z=z01 ‡ A2 An0 exp z=z02 dz dliq
dliq z


dliq z
‡  An0 : …114†
dliq

De®ning cp as the integral over the exponentials

X2     
Ai z0i z dliq
cp ˆ exp exp ; …115†
d
iˆ1 liq
z0i z0i

7
Here and in the following, a number without the star means irradiation in natural nitrogen, whereas a
number with a star means irradiation in the enriched atmosphere.
76 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Eq. (113) can be written as


2 dliq
!3 i
z
z0 z0
n 16
z02 e 2 e 2
7
X 6dliq z 7
An‡1 ˆ cp  6 7: …116†
0 6 dliq dliq 7
iˆ0 4 5

With the values given above we obtain cp ˆ 0:32 at.% and the o€set An0 can now be
calculated for all n and the development of the nitrogen depth pro®les with n is dis-
played in Fig. 49. The simulation is in excellent agreement with the experimental
results and properly explains the observed saturation in the nitrogen depth pro®le.
This is a strong support for the simple scenario since the parameters are determined
independently.
The saturation behavior during the laser nitriding of iron is caused by the com-
petition between the take-up of new nitrogen, the ablation of the surface and the
lateral transport of material out of the center of the laser spot. This and the
deposition of fall-out can be easily revealed by exploiting the nuclear reaction
15
N…p; †12 C for nitrogen pro®ling and carrying out the laser treatments in nitro-
gen atmospheres with alternating 15 N enrichments. Thus, the isotopic sensitivity of
RNRA is a valuable tool to resolve the complicated and superimposed e€ects during
the laser treatment of iron.
The simple theoretical description given here is a consequence of the strong lateral
material transport due to the piston mechanism. This leads to very low nitrogen
concentrations before each subsequent pulse, so that the di€usion and take-up
mechanisms are not changed by high nitrogen concentrations already present in the

Fig. 49. Simulation of the nitrogen depth pro®les. The solid lines are drawn according to the presented
model with the mean values given above. The insert shows the saturation behavior and the solid line is
drawn according to Eq. (116).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 77

surface layer. If the nitrogen concentration would be higher, signi®cant changes are
expected and the modeling would become more complicated.
The validity of the model can now be cross-checked by the reverse experiment, i.e.
by ®rst irradiating in an enriched atmosphere followed by irradiations in natural
nitrogen. The resulting nitrogen depth pro®les as measured with RNRA are dis-
played in Fig. 50. The 15 N depth pro®les can again be ®tted with Eq. (110) and the
resulting parameters for sample (64/0) are: c1 ˆ 18…2† at.%, z1 ˆ 14…4† nm, c2 ˆ
15…2† at.% and z2 ˆ 92…9† nm. These are in agreement with the mean values given
above. The o€set amounts to c0 ˆ 3:7…2† at.%. For sample (63/1) it becomes
already obvious that the take-up of new 15 N with the 64th pulse, carried out only in
natural nitrogen, is missing and only the nitrogen loss during this last pulse is
observed. This is even more pronounced for sample (62/2). The ®ts according to
Eq. (110), for which the parameters z1 and z2 were constrained to the values for
sample (64/0), give the following parameters for sample (63/1): c0 ˆ 3:6…2† at.%,
c1 ˆ 0:4…2† at.%, and c2 ˆ 5:3…4† at.%, and for sample (62/2): c0 ˆ 3:1…2† at.%,
c1 ˆ 0:4…3† at.%, and c2 ˆ 0:4…3† at.%.
The parameter c1 (fall-out) nicely follows the factor 45 of the enrichment, i.e. it
resembles the concentration of 15 N in the treatment gas. Parameter c2 for sample
(63/1) is much higher than expected from the enrichment factor. This may be
explained by the ablation and re-dissolution of 15 N present with high concentration
in the surface of the treated material. Its decrease to the expected value from sample
(63/1) to sample (62/2) supports this explanation. According to the model pre-
sented in [45], the new o€set can be calculated via Eq. (113). When using the para-
meters given above, the surface loss z = 55 nm and the thickness of the liquid
surface dliq ˆ 800 nm, the following o€sets are obtained: c0 ˆ 3:7 at.% for sample

Fig. 50. 15 N depth pro®les obtained for 64 pulses in enriched atmosphere (64/0), 63 pulses in enriched
atmosphere followed by the 64th pulse in natural nitrogen (63/1), and 62 pulses in enriched nitrogen
followed by the 63th and 64th pulse in natural nitrogen (62/2) (all at 4 J/cm2, 1013 hPa).
78 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

(63/1) and c0 ˆ 3:4 at.% for sample (62/2). This is in good agreement with the
experimental values.
The bene®ts of these alternating irradiations in natural and enriched atmospheres
can also be used for the measurements of the lateral nitrogen concentrations across
the laser spot. They were obtained by scanning the analyzing proton beam (2 mm
diameter) at the ®xed proton energy of Ep ˆ 432:5 keV (z  20 nm) across the sur-
face, i.e. at a ®xed analyzing depth. Fig. 51 compares the nitrogen contents for a
sample irradiated with 64 pulses in enriched atmosphere (64/0) with a samples
irradiated with 63 pulses in natural nitrogen followed by a single pulse in enriched
atmosphere (63/1) and a sample irradiated with 63 pulses in enriched atmosphere
followed by one pulse in natural nitrogen (63/1).
The characteristic two-peak structure was explained by the fall-out having a high
nitrogen content of about 30 at.% and can be well approximated by the exponential
decay of its thickness with the distance from the edge of the laser spot [55].
Obviously, sample (64/0) can be seen as the superposition of samples (63/1) and
(63/1). Sample (63/1) does not show the fall-out because there the fall-out only
consists of 14 N and only the Gaussian peak of the newly incorporated 15 N is visible
with RNRA. The latter is about the same as obtained for a single pulse in 15 N. In
analogy, sample (63/1) only shows the fall-out consisting of the 15 N already present
in the sample, whereas the new nitrogen coming in with the 64th pulse in natural
nitrogen is invisible for RNRA.
At a greater distance from the laser spot, the previously mentioned exponential
decay of the fall-out thickness dfo can also be observed when scanning the lateral
nitrogen pro®les with RNRA and increasing the proton energies. Fig. 52 shows
these lateral nitrogen pro®les for sample (63/1), taken with increasing proton ener-
gies Ep . As shown before, sample (63/1) mainly shows the fall-out, not the take-up
of new nitrogen. The constant nitrogen concentration in the center of the laser spot

Fig. 51. Lateral nitrogen pro®les as obtained by RNRA for samples: (a) (0/1), (b) (63/1), (c) (64/0) with
Ep ˆ 432:5 keV (18 nm) (the bar indicates the lateral size of the laser spot).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 79

for all energies, i.e. for all depths, again shows the constant nitrogen depth pro®le
for this sample (compare with Fig. 50). Furthermore, as shown in Fig. 52, the height
and the width of the peaks are decreasing with increasing proton energy, i.e. with
increasing detection depth. This is in full agreement with the exponential decay of
the fall-out thickness as measured before with surface pro®lometry [50,55]. A ver-
i®cation of this is displayed in Fig. 53, where the total amount of the fall-out is dis-
played against the proton energy, i.e. against the measuring depth. A good
agreement with the exponential decay of the fall-out thickness can be and the decay
constant is 32(4) keV. This can be transformed into a decay depth of 208(26) nm,

Fig. 52. Lateral nitrogen pro®les for sample (63/1) obtained at various proton energies.

Fig. 53. Integrated amount of fall-out versus proton energy (depth).


80 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

which corresponds to a fall out deposition of 3(1) nm/pulse, in excellent agreement


to the value of 4 nm/pulse, estimated from the surface pro®lometry [55,50].
The nitrogen di€usion and transport during each following laser pulse can also be
observed in Fig. 54, where the irradiation with one pulse in enriched atmosphere is
followed by 0, 1, 2, and 4 pulses in natural nitrogen. The nitrogen depth pro®les are
almost immediately homogenized with the exception of a small surface peak.

6.6. In¯uence of the energy density

In Section 4.3 it was estimated that signi®cant evaporation (surface boiling) and
thus plasma formation starts at about 2.4 J/cm2. Of course this result has to be
con®rmed experimentally. Therefore, nitrogen depth pro®les were measured for
samples irradiated with various average laser ¯uences ranging from 1 to 4 J/cm2 and
for 64 pulses. The ¯uence was set by adjusting a variable beam attenuator. These
depth pro®les show that there is already a nitriding e€ect for H ˆ 1:5 J/cm2. By
increasing the energy density the nitrogen concentration at the surface is increasing,
but decreases in deeper regions. As a measure of the nitriding e€ect the mean
nitrogen concentration averaged over the topmost 300 nm8 is displayed in Fig. 55. A
threshold energy density of about 2.0(2) J/cm2 was given for melting and plasma
formation induced by excimer laser irradiation of iron [316,28]. Here we ®nd that
there is a sharp onset of the nitriding e€ect already at about 1.5 J/cm2, which then
remains more or less constant. For an explanation of this ®nding one should keep in
mind the following two facts: (i) the ¯uence is given as a mean value over the whole
laser spot and the spatial intensity distribution is quite inhomogeneous; the laser
intensity is highest in the center of the spot and lies about 50% above the mean (see

15
Fig. 54. N depth pro®les at 4 J/cm2: (1/0), (1/1), (1/2) and (1/4).

8
This range is easily accessible with RNRA at the IONAS
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 81

Fig. 19), (ii) the depth pro®les are taken in the center of the spot, and the observed
lateral transport may be di€erent for the di€erent energy densities used.
Indeed these e€ects are con®rmed in the lateral nitrogen pro®les shown in Fig. 56.
Here the energy density was varied by setting the laser spot area. For 1.2 J/cm2 only
a small region in the central part is nitrided and the concentration is still low.
Increasing the laser ¯uence increases the central nitrogen concentration and the
width of the nitrided region. At 2.2 J/cm2 we have the broadest nitrided region, there
is a nitriding e€ect over the whole laser spot and the lateral material transport
(piston e€ect) starts. At 4 J/cm2, we found a strong `piston' e€ect and the nitrided is

Fig. 55. Mean nitrogen content (300 nm) as a function of the energy density H for the raw beam (64 pulses).

Fig. 56. Lateral nitrogen pro®les for 32 pulses at various energy densities measured at Ep ˆ 432:5 keV
(20 nm).
82 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

region smaller again. Thus the threshold average laser ¯uence for the nitriding
should be around 1.2 J/cm2. The maximum concentration in the lateral peaks is
independent of the laser ¯uence (if above the nitriding threshold), as expected from
their origin in the nitride fall-out. The dip in the central part becomes deeper with
the ¯uence H, due to the enforced lateral transport.

6.7. In¯uence of the spot size

In the experiments described before, the energy density of the laser was set by a
variable attenuator or by varying the size of the laser spot at the sample surface. The
latter changes the size of the laser spot and the laser intensity simultaneously. In the
next step the in¯uence of both parameters was studied independently. First the spot
size was changed by setting the laser energy density constant to 2.5 J/cm2 by proper
adjustment of the attenuator for the spot size investigated. Then for a constant spot
size of Ap ˆ 14:6 mm2 by changing the mean laser ¯uence H through the attenuator.
All these experiments are carried out for 64 pulses in natural nitrogen.
According to the theory of the pistonpmechanism
 [272] the lowering zpist should
increase with the laser spot area like 1= Ap (see Eq. (53)) and with the laser inten-
sity (see Eq. (40)). The surface pro®les shown in Fig. 57 hardly change with the
energy density and the piston e€ect does not seem to become more important for
H ˆ 4, 6, and 8 J/cm2, even the two peak structure on the left borders is present in
all cases. The amount of fall-out is increasing with the energy density H and also
spreads further away from the spot as was veri®ed by optical inspection. A numer-
ical calculation of the total transported material should clarify the e€ect of changing
the laser ¯uence. The e€ect can be described again by d‡ and d (see Eq. (109)),
i.e. the material moved from the center towards the borders above the surface, i.e.
from the trough to the ridges.

Fig. 57. Surface pro®les of iron irradiated with a laser spot of Ap ˆ 14:6 mm2 at (a) 4 J/cm2, (b) 6 J/cm2
and (c) 8 J/cm2 (64 pulses) (a and b are shifted by 30 and 15 mm for better visibility).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 83

The amounts of transported material obtained from the pro®les shown in Fig. 57
are displayed in Fig. 58. There is no obvious relationship between the amount of
moved material and the laser ¯uence. However, the di€erence d ˆ d ‡ d ,
calculated here to be d ˆ 30…6† nm/pulse, is not necessarily related to a real mate-
rial loss, because this is only the two-dimensional value in the center of the spot (see
Section 6.1). Nevertheless, the laser ¯uence is not the most important parameter for
the piston e€ect, at least in the measured parameter range.
Fig. 59 shows the surface pro®les of iron irradiated with 2.5 J/cm2 but di€erent
laser spot sizes. It is clearly seen, that the piston e€ect, i.e. the depth of the trough
and the height of the ridges, is signi®cantly increased when the spot size decreases.

Fig. 58. Material transport for increasing laser ¯uence at unchanged laser spot area.

Fig. 59. Pro®lometer measurements for H ˆ 2:5 J/cm2 and 64 pulses with varying spot size Ap (a and b
are shifted by 8 and 4 mm for visibility).
84 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

This is expected since Eq. (53) gives z / A 1=2 , and the surface pressure should
remain unchanged due to the constant laser ¯uence.
The amount of transported material is presented in Fig. 60. Surprisingly, a beha-
viour d  / A 1 is observed, which is unexpected from the piston theory. In order
to resolve this di€erent behavior by changing the spot size and the laser ¯uence
alone, one has to look on the beam optics of the irradiation treatment (see Fig. 20).
If there is a inhomogenity (¯uctuation) in the spatial intensity distribution of the
laser beam, this will be ampli®ed by focusing the beam. Fig. 61 shows an illustration
for this fact. It is assumed that the ¯uence has a triangular distribution and that the

Fig. 60. Material transport against the laser spot area for 2.5 J/cm2.

Fig. 61. Increase of inhomogenities by focusing. The lateral form of the intensity distribution is triangular.
For the unfocussed original pro®le it is assumed that Hc ˆ 1:5  Hb .
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 85

¯uence in the central part Hc is 50% higher than the intensity at the border of the
laser spot Hb : Hc ˆ …1 ‡ i†  Hb with i = 0.5, and the focusing is given in multiples of
the original mean laser ¯uence. Recalling the spatial pro®le of the raw beam (see
Fig. 19) the value of 50% seems adequate. The tip is getting sharper for enhanced
focusing as expected. The increase in the inhomogenity of the laser ¯uence changes
with the spot size Ap ˆ A0  f 2 d2 (see Section 5) according to

A0
H ˆ Hc Hb ˆ i  Hb  …117†
Ap

and the gradient is given by


 3=2
H A0
ˆ i  Hb  : …118†
x Ap

The assumption of a linear relationship of pressure and ¯uence (pr / H and


pp / H2=3 ) leads to z / A 1 . The surface lowering z is displayed in Fig. 62 for all
the previous shown areas and laser ¯uences and the best ®t is obtained for
z / A 3=2 .
This is a good indication, that the pressure gradient inside the laser spot is
responsible for the lateral material transport. The pressure gradient would also
induce a stronger plasma expansion and thus would explain the further distribution
of the fall-out outside the spot area for smaller spot sizes.
Nevertheless, for a full understanding of this lateral material movement, the
complete hydrodynamical equations have to be solved for a given pressure dis-
tribution above the surface, where this pressure distribution itself has to be obtained
from simulations of the laser plasma behavior.

Fig. 62. Mean lowering z in the center of the laser spot (averaged over 1 mm  x  ‡1 mm), the solid
lines are ®ts to z ˆ c=A3=2
p .
86 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Due to the importance of the laser spot size on the lateral material transport, also
the nitrogen take-up depends on it. Fig. 63 shows the 15 N depth pro®les obtained
after irradiations with 4 J/cm2 and n pulses in natural nitrogen followed by one pulse
in enriched 15 N. The laser spot size was increased from 17.9 to 27.9 mm2 and thus
the piston e€ect is reduced. The change in the nitrogen depth pro®les now clearly
indicate that the take-up of new nitrogen is dependent on n. This is explained by the
drastically reduced piston e€ect and thus reduced material transport leading to
higher nitrogen concentrations already present in the surface.
The results of the ®ts according to Eq. (110) for this pro®les show that all para-
meters except z2 are remaining constant for increasing n. z2 is decreasing from about
90 nm for n ˆ 1 to about 45 nm after 512 pulses. As seen before the larger spot size
decreases the lateral transport and higher nitrogen concentrations are achieved in
the center of the spot. This lowers the concentration gradients and the di€usion
length is reduced with the number of pulses.

6.8. Microhardness measurements

In this section the overall surface hardness and hardness pro®les of the nitrided
layers after the raw beam irradiation will be discussed. Fig. 64 shows the lateral
surface hardness of iron irradiated with 8 pulses at 4 J/cm2 in 1013 hPa nitrogen gas.
The hardness is measured with a load of 4 mN (HU 0.004), so that the indentation
depth is in the order of 100±150 nm, thus measuring the hardness of the nitrided
surface layer [493]. An optical view of the laser irradiated spot is added for
comparison.
The increase in hardness is well correlated with the optically visible laser spot. The
increase is more or less constant across the spot and no signs of the piston e€ect can
be observed in the lateral hardness. The development of this surface hardness with

Fig. 63. Nitrogen depth pro®les for (n/1) with larger spot size (Ap ˆ 27:9 mm2, 4 J/cm2).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 87

the number of laser pulses is displayed in Fig. 65. The nitride fall-out is not observed
in the hardness pro®le. It is important to note that the hardness of the e-nitride is
about 7.1(8) GPa [494] and agrees with the hardness inside the laser spots. When
comparing the lateral hardness of the raw beam sample with the lateral surface
pro®le (see Fig. 70), it becomes evident that the whole laser spot shows homo-
geneous and high hardness values and does not follow the surface pro®le showing
deep craters and high ridges.
The surface hardness is increasing with the number of pulses and seems to become
also more homogeneous across the laser spot. There is always a sharp rise or decay
of the hardness at the border of the spot. Averaging over all values inside the spots

Fig. 64. Lateral scan of the surface hardness (HU 0.004) across a laser irradiated spot (8 pulses at 4 J/cm2
in 1013 hPa nitrogen). An optical view of the spot is included.

Fig. 65. Lateral scans of the microhardness across the laser spots for various n (4 J/cm2).
88 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

gives the mean values as displayed in Fig. 66. The same saturation behavior with the
number of pulses is observed in the hardness as already seen in the nitrogen depth
pro®les. The hardness given in GPa increases with the number of pulses according to
M ˆ 2:5…3† ‡ 4:5…3†  …1 exp… n=6:5…8†††. The saturation is reached at about 32
pulses and the saturation value is 7.0(4) GPa, typical for the e-nitride.
The dependence of the surface hardness on the laser ¯uence was measured for the
same samples as scanned by RNRA (see Fig. 55). Also here the hardness starts to
increase already at about 1.2 J/cm2 and saturates for ¯uences above 2 J/cm2, as seen
in Fig. 67.

Fig. 66. Mean surface hardness as a function of n (4 J/cm2, 1013 hPa, there is an additional point at
n = 512).

Fig. 67. Mean surface hardness as a function of H (32 pulses).


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 89

In Fig. 68 the variation of the hardness depth pro®les with the pulse number is
displayed. Apart from the region next to the surface the pro®les qualitatively
resemble the nitrogen concentration pro®les, with high hardness values at the sur-
face which then decrease when going to greater depths. As in the case of the nitrogen
take-up and the surface hardness, the increase of the hardness saturates after 32
pulses for irradiation with the raw beam. The hardening depth is roughly 2 mm,
where the hardness of the untreated and the irradiated sample merges.
The hardness depth pro®les as function of the laser ¯uence, as given in Fig. 69,
again show a similar behavior as found in the previous measurements. The hard-
ening starts already at 1.2 J/cm2 and saturates after 2.5 J/cm2.

Fig. 68. Hardness depth pro®les for increasing pulse numbers at 4 J/cm2.

Fig. 69. Hardness depth pro®les for 32 pulses at various energy densities.
90 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

6.9. Summary of raw beam irradiation

The saturation behavior during the laser nitriding of iron is caused by the com-
petition between the take-up of new nitrogen, the evaporation/ablation of the sur-
face and the lateral transport of material out of the center of the laser spot (piston).
This and the generation of fall-out can be easily demonstrated by exploiting the
nuclear reaction for nitrogen pro®ling and carrying out the laser treatments in
nitrogen atmospheres with alternating enrichments in the isotope 15 N. Thus, the
isotope sensitivity of RNRA is a valuable tool to resolve complicated and super-
posed e€ects during the laser treatment of surfaces.
The lateral nitrogen, surface and hardness pro®les are again compared in Fig. 70.
The lateral hardness increase is restricted inside the ridges formed due to the piston
e€ect. The surface pro®le is completely located inside the spot, whereas the maxima
of the lateral nitrogen distribution are lying outside the laser spot. The lateral
expansion of the plasma deposits the re-condensing nitrides also outside the laser
spot.
The lateral transport is caused by pressure gradients induced by the inhomoge-
neous raw laser beam and is drastically depending on the laser spot size and only
marginally on the average laser intensity. A modeling of the saturation behavior is
possible for series with large lateral material transport, because there only low
nitrogen concentrations are present in the center after each pulse, not hindering the
new nitrogen take-up. Less lateral transport leads to higher nitrogen concentrations
in the center of the spot and thus the take-up of new nitrogen starts to decrease and
to saturate after a certain number of pulses.
There is also a strong convection in the liquid state, intermixing and homogeniz-
ing the concentrations. The convection is probably stronger for larger gradients in
the laser intensity and smaller laser spots. The hardness is increased by a factor of
three reaching values of about 7 GPa. The hardening depth is about 2 mm and

Fig. 70. Comparison of all lateral measurements (RNRA, pro®lometer, hardness).


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 91

coincides with the calculated and observed thickness of the molten layer. The
resulting phase formation after the nitriding treatment will be discussed in Section 8.

7. Homogenized beam irradiations

The spatial intensity distribution of the raw laser beam used in Section 6 is far
away from a ¯at top-hat pro®le (see Fig. 19). One might ask which e€ects occurring
during the laser nitriding process are related to the inhomogeneous spatial pro®le of
the laser beam. Therefore, experiments have been carried out with an additionally
homogenized laser beam.
An optical view (topview) of a laser irradiated spot using the homogenized laser
pulse is shown in Fig. 71. It becomes immediately clear that the e€ects of the
homogenized beam drastically di€er from those observed for the raw beam irradia-
tions shown in Fig. 34. There is a sharp border of the laser spot, no inhomogenities
inside the spot and almost no e€ects outside the spot are visible. Thus, the homo-
geneous beam pro®le leads to a much more homogeneous spot structure. The pro-
nounced periodicity inside the laser spot found for the raw beam irradiation is less
pronounced here for the homogenized beam. Nevertheless, some surface waves are
also visible, as proven in Fig. 72. But the waves, also related to the plasma above the
liquid surface, seem to be smoother. The periodicity is also about 30 mm and the
corrugations are of the order of 500 nm in the depth. The large amount of fall-out
found outside the laser spot for the raw beam irradiations is completely missing for
the homogenized beam irradiations, only for 512 pulses, some small spherical par-
ticles are found outside the laser spot.

Fig. 71. Optical view of the homogenized laser spot (4 J/cm2, homogenized beam, 64 pulses).
92 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

7.1. Surface pro®les

The surface topology induced by the homogenized beam, as already seen in the
optical pictures before, was further investigated with the pro®lometer. Fig. 73 shows
the surface pro®les obtained after irradiation with various numbers of pulses of the
homogenized beam. For low n, the surface remains much smoother as compared to
the raw beam irradiations (see Fig. 37). Nevertheless, also here an increase in the
surface roughness and a `piston' e€ect is visible, but less pronounced. The ridges
of the laser induced crater are clearly present for n ˆ 256 and 512. Some surface
corrugations also exist outside the laser spot for the higher number of pulses.
The normalized layer thickness of the material transport (d ‡ and d ), again
obtained by integration of the positive and negative parts of the surface pro®le,
shows again a linear relation with the number of pulses n, as displayed in Fig. 74.
The calculated slopes are 0.5(1) and 1.4(2) nm/pulse. These values are about 20
times smaller than those of the raw beam irradiations.
Again, the surface lowering z in the center of the spot was calculated as the
mean of the surface height over the range 1 mm. The values obtained from the
surface pro®les are given in Fig. 75. Linear regression gives z ˆ 1:8…4† nm/pulse.
This value is smaller than the ablated layer thickness of 8 nm/pulse, determined by
the simulation in Section 4.3. This indicates that most of the ablated material must
be again re-deposited inside the laser spot. The total loss is estimated to be 0.4(3)
nm/pulse, much smaller than the value for the raw beam irradiation (4(2) nm/pulse).
These `smooth' results of the homogenized beam irradiations can be veri®ed in the
SEM picture of a brittle fracture in Fig. 76. A ¯at surface with a ¯at laser in¯uenced
layer is seen there. No lateral inhomogenities are found and the thickness of the laser
a€ected layer is 750(100) nm. This agrees excellently with the 800 nm molten layer
thickness derived from the simulation (see Section 4.3).

Fig. 72. Surface waves inside the homogenized laser spot (4 J/cm2, 64 pulses).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 93

Fig. 73. Surface pro®les for various numbers of pulses (4 J/cm2 homogenized beam). The two vertical
lines are marking the laser spot dimension.

Fig. 74. Material transport for homogenized irradiations (4 J/cm2, homogenized beam).
94 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

The increase of the surface roughness Ra is displayed in Fig. 77 and it shows the
same saturation with n as the laser nitriding with the raw beam.
p When ®tting the
data points with the square root dependency Ra ˆ R0a ‡ R0a  n, the value R0a ˆ
15…2† nm is obtained.

7.2. Lateral nitrogen pro®les

The drastically reduced `piston' e€ect should also have an in¯uence on the lateral
nitrogen distribution. Fig. 78 shows the lateral nitrogen pro®les for various numbers
of pulses as obtained by RNRA measurements. The pro®les are taken with a proton
energy of 432.5 keV which analyzes a depth of about 20 nm. The nitrogen pro®les

Fig. 75. Mean surface lowering z in the center of the homogenized spot.

Fig. 76. Brittle fracture of a homogenized irradiation as observed in the SEM (4 J/cm2, 64 pulses).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 95

are laterally homogeneous. There is no double peak structure evolving at the edges
as was observed for the raw beam. The nitrogen concentration in the center is
steadily increasing with n from 2 at.% for a single homogenized pulse to a saturation
value of about 25 at.% for 128 or more pulses. At the border of the laser spot,
indicated by the bar, the nitrogen concentration rapidly falls to zero and no nitrogen
can be found outside the laser spot. This proves that no `fall-out', containing nitro-
gen, is present outside the laser spot. At least the amount of fall-out or its con-
centration is too low to be detected (  0:5 at.%).

Fig. 77. Surface roughness for the homogenized irradiations.

Fig. 78. Lateral nitrogen pro®les in dependence of the number of pulses n (4 J/cm2, homogenized beam)
(analyzing beam: Ep ˆ 432:5 keV, 1 mm diameter).
96 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

7.3. Nitrogen depth pro®les

The di€erence of the lateral nitrogen pro®les as compared to the raw beam irra-
diation may be also re¯ected in the depth distribution. Fig. 79 shows the nitrogen
depth pro®les as obtained from the RNRA measurements in the center of the laser
spot. For a single homogenized laser pulse, one ®nds only about 3 at.% nitrogen at
the surface and also a fast decrease with depth. For an increasing number of pulses
the surface concentration develops to a saturation value of about 30 at.%. In greater
depths, also more nitrogen is found until the depth pro®le saturates for 128 and
more pulses.
This is also re¯ected in the total amount of nitrogen found in the ®rst 300 nm,
given as the mean nitrogen concentration hcN i. A steep increase and then a satura-
tion behavior is observed. The values hcN imax ˆ 16:6…5† at.% and n0 ˆ 31…2† are
obtained when ®tting the data with the activation function hcN i ˆ hcN imax  …1
exp… n=n0 † ).
In analogy to the raw beam, these depth pro®les can be well ®tted according to
Eq. (110) and the results are displayed in Fig. 80. The saturation is seen in the
behavior of the o€set c0 , the saturation value is 15.0(6) at.% with n0 ˆ 62…9† pulses.
The other parameters c1 , z1 , c2 , z2 need some 10±20 pulses until they reach a con-
stant value and this can be described using an activation behavior again. c1 is
decreasing from more than 30 at.% to the value 7.1(4) at.%, whereas c2 is increasing
from about 2 at.% to the saturation value of 20(1) at.%. The di€usion length z1 is
increasing from almost zero to 33(2) nm, whereas z2 is increasing from about 100 nm
to the saturation value 209(8) nm. The mean activation `time' is 16(8) pulses for all
parameters.
As in the case of the raw beam irradiations a short and a long range di€usion is
found, and again they are attributed to the di€usion of nitrogen (c2 , z2 ) and the fall-

Fig. 79. Nitrogen depth pro®les for increasing number of pulses n (homogenized beam, 4 J/cm2) (n  128
are shown as lines only).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 97

out deposition (c1 , z1 ). The adjustment of the parameters before they reach their
constant values may be due to incubation e€ects. Maybe, the amount of evaporated
material is increased or the plasma ignition is easier, when some roughness is
induced by the ®rst pulses, so that tips in the surface allow an easier evaporation of
material and thus favor a plasma formation.

7.4. Irradiation of marker layers

The material transport and the intermixing in the melt was again examined using
Au marker experiments. Fig. 81 shows the RBS spectra of a Au marker layer in a
depth of about 90 nm and a thickness of about 10 nm before and after irradiation
with the homogenized laser beam. The Au concentration pro®les as obtained from
the RBS spectra are shown. The Au marker layer is clearly visible before the laser
treatment. After a single laser pulse there is still a very broad Gaussian Au pro®le
visible, which moved towards the surface. After two and more pulses a homo-
geneous Au-pro®le is achieved. The position of the Au marker layer is shifted from
92 to 52 nm and at the same time its width is increasing from 7 to 36 nm. After two
or more pulses the concentration is constant. This can be interpreted in the follow-
ing way: there is no complete intermixing during the irradiation with a homogenized

Fig. 80. Parameters for the ®ts of the nitrogen depth pro®les (4 J/cm2, homogenized beam). Fits are
performed using Eq. (110).
98 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

laser pulse. There is less convection than compared to the raw beam, but there still
is. Even the material transport is smaller for the homogenized beam irradiation, the
shift of the Au-peak towards the surface indicates that still some macroscopic
material transport is e€ective.
For the case of pulsed laser alloying it is known that the spatial laser intensity
distribution determines the intensity of melt convection and therefore the distribu-
tion of the alloying elements. For good convective mixing an inhomogeneous
intensity distribution is favorable [495]. Accordingly, the transport of nitrogen into
the sample is slower for irradiations with the homogeneous beam. On the basis of
the experiments carried out with the Au marker layers located at various depths z0
from the iron surface, the degree of intermixing in the liquid can be distinguished for
raw beam and homogenized beam irradiation. The Au concentrations measured by
RBS at the surface of the various samples are displayed in Fig. 82. The steady
decrease of the Au surface concentration under homogeneous irradiation shows that
there is no complete intermixing but that the Au mainly di€uses towards the surface.
Contrary to this, the Au surface concentration does not really depend on the depth
of the marker layer under raw beam irradiation, indicating a strong convection
during the pulse over the whole liquid depth.
The piston e€ect and the material loss in the center of the laser spots can be seen
in Fig. 83, which shows the total Au content as measured by RBS. For the homo-
geneous beam, the Au content only slightly drops with the number of pulses,
whereas the raw beam induces considerable losses of Au out of the center, even for
marker layers in large depths (z0 ˆ 99 nm). The strong lateral material transport for
the raw beam leads to a faster saturation of all parameters as compared to the
homogenized beam.

Fig. 81. Irradiation of Au marker layers with the homogenized beam: (a) RBS spectra of a Au marker
layer before and after laser irradiation; (b) Au concentration pro®les as calculated from the RBS spectra.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 99

7.5. Isotopic experiments

The advantages of the combination of the RNRA method with laser treatments in
isotopic enriched atmospheres was exploited also for the homogenized beam.
Fig. 84 shows the 15 N depth pro®les obtained by the irradiation with n pulses in
natural nitrogen followed by one pulse in the enriched nitrogen atmosphere (n/1). It
becomes clear that the take-up of new nitrogen visible here is depending on the his-
tory of the sample, i.e. the number of previous pulses shot at the surface. The surface
concentration is increasing with n, whereas the depth where new nitrogen is found is
decreasing with n.

Fig. 82. Au surface concentration as function of the marker depth (n is given).

Fig. 83. Normalized total Au content in the center of the laser spot as function of n (z0 is given).
100 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

This becomes clearer after ®tting the depth pro®les according to Eq. (110) and the
results are displayed in Fig. 85. We ®nd c0 ˆ 0, since there is no 15 N present in the
sample before. The other parameters are more or less constant and only z2 exhibits a
clear dependence on n, its value decreasing from about 100 nm to about 60 nm. As
seen before in Fig. 80, as much as 15 at.% nitrogen can already be present in the
sample. The lowering of z2 is in accordance with the increasing nitrogen content
already present in the sample, thus lowering the concentration gradient and along
with that the amount of nitrogen di€used into the surface layer is decreased.

Fig. 84. 15 N depth pro®les for homogenized beam irradiation with n pulses in natural nitrogen followed
by 1 pulse in enriched nitrogen (n/1).

Fig. 85. Results of the ®ts to the nitrogen depth pro®les …n=1 †.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 101

The results of the reverse experiment, irradiation in enriched atmosphere followed


by irradiation in natural nitrogen (…64 n† =n†, is shown in Fig. 86. The sample
(64 =0) of course exhibits the normal nitrogen depth pro®le. Sample (63 =1) shows a
large decrease of the nitrogen concentration at the surface, whereas in deeper layers
the concentration remains unchanged as compared to sample (64 =0). The fall-out
contribution to the depth pro®le disappeared. The parameter c2 decreased from 8
at.% for sample (64/0) to 6 at.% for sample (63/1), and down to 5 and 2 at.% for
samples (62/1) and (60/4), respectively. Due to the small piston e€ect, not much
15
N should be lost with the subsequent pulses. A di€usive and possibly convective
distribution of the nitrogen should be active, which destroys the high surface con-
centrations. So far the model is con®rmed since the o€set c0 is not decreasing with n,
but remains constant at c0 ˆ 8…1† at.%.

7.6. Modeling of nitrogen depth pro®les

With the knowledge collected so far it was attempted to simulate the nitrogen
depth pro®les for the homogenized beam irradiations. Therefore, the simulation
program used for the temperature pro®les was extended to calculate the nitrogen
concentration pro®les, using the same equations and principles of the method of
®nite di€erences. Only new arrays of concentration and di€usion constants had to
be implemented. The di€usion was restricted to the liquid state and the di€usion
coecient was assumed to be temperature dependent according to Eq. (9) but
neglecting its concentration dependence. Furthermore, it was assumed that the take-
up of new nitrogen is limited to the laser pulse duration (55 ns) and afterwards only
di€usion of nitrogen takes place. The nitrogen dissolves in the topmost layer with a
solubility of c2 ˆ 20 at.%. The amount of fall-out (c1 , z1 ) deposited onto the surface
after the laser pulse was also taken from the results of the depth pro®le ®ts. The

Fig. 86. Nitrogen depth pro®les obtained for ……64 n† =n†.
102 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

results of this simulation are compared in Fig. 87 with the measured depth pro®les.
Although there are di€erences in the absolute values, the qualitative behavior of the
development of the depth pro®les with the number of pulses is nicely reproduced.
Thus, di€usion can explain the development of the nitrogen pro®les for the homo-
genized beam. Nevertheless, there is also some convective transport present (as seen
in the Au marker experiments), not included in the simulation.
Additionally, the depth pro®les for isotopic irradiations are also reproduced
rather well as proven in Fig. 88. There, the depth pro®les obtained after irradiation
with a single pulse in the enriched atmosphere followed by 0±4 pulses in natural

Fig. 87. Nitrogen depth pro®les and simulation (see text).

Fig. 88. Nitrogen depth pro®les of isotopic irradiations …1 =n†.


P. Schaaf / Progress in Materials Science 47 (2002) 1±161 103

nitrogen (1 =n) are displayed. Again the simulation reproduces the depth pro®les
rather well.

7.7. In¯uence of the energy density

Since the irradiations were carried out with a spatially very homogeneous laser
beam, it was of special interest to investigate the in¯uence of the energy density on
the nitriding process. Unfortunately, for this kind of experiments is was not possible
to achieve ¯uences above 4 J/cm2. Fig. 89 shows the lateral nitrogen pro®les as
obtained by means of RNRA after laser nitriding at H ˆ 0:5 4:0 J/cm2. For ¯u-
ences below 2 J/cm2 there is hardly any nitriding e€ect. For energy densities of 2.5 J/cm2
and above nitriding can be clearly observed. The width of the lateral nitrogen dis-
tribution is increasing with the energy density, although the laser spot size is
unchanged.
The similar behaviour is observed in the nitrogen depth pro®les shown in Fig. 90.
Below 2.3 J/cm2 there is little nitriding. At 2.5 J/cm2 a high nitriding eciency starts.
The pro®le itself is unchanged for increasing energy densities, but shifted towards
higher concentrations. The parameter z2 is constant, irrespective of the energy den-
sity, thus the di€usion time is not in¯uenced by H.
This dependence of the nitriding eciency on the laser ¯uence is plotted in Fig. 91.
The mean nitrogen concentration in the topmost 300 nm hcN i shows a slow (expo-
nential) increase with the energy density up to 2.5 J/cm2. Just above the theoretically
calculated threshold of 2.4 J/cm2, there is a sharp increase of the nitrogen take-up. The
mean nitrogen concentration then still increases with the laser ¯uence up to 4 J/cm2.
Due to the limitation in laser power it was not possible to reach a saturation. The
threshold for the strong nitriding e€ect is 2.4(2) J/cm2, although already at 2 J/cm2
nitriding is enhanced above the normal solubility.

Fig. 89. Lateral nitrogen pro®les as function of the laser ¯uence (homogenized, 64 pulses, Ep ˆ 432:5
keV).
104 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

The depth pro®les were again ®tted with di€usion pro®les. It was found that the
energy density dependence of the nitriding eciency is mainly expressed by the
parameter c2 , which is also shown in Fig. 91. The same behaviour as for hcN i is
observed. Up to the threshold, c2 can be explained with the solubility of nitrogen in
liquid iron (or g-Fe) at atmospheric pressure (Eq. (1)). The temperatures for this
calculation were taken from the simulations. The values for 2 and 2.3 J/cm2 are
already somewhat enhanced, due to the starting plasma formation. Above threshold
the high nitrogen concentration can only be explained by the high pressures induced

Fig. 90. Nitrogen depth pro®les for various energy densities (homogenized, 64 pulses).

Fig. 91. Mean nitrogen concentration in the surface as derived from the depth pro®les. The surface con-
centration c2 as obtained from the ®ts of the depth pro®les is also shown. The dashed line is calculated
with the nitrogen solubility (see text).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 105

by the LSC waves in addition to the recoil pressure. Eq. (1) was used again together
with the plasma pressure (LSC) and an enhanced solubility due to the plasma for-
mation (here the factor 3.4 was used, [156]). It is found that this pressure induced
high solubility can explain the observed values for c2 , as indicated by the dashed line
in Fig. 91.

7.8. Microhardness measurements

As the surface topography and nitrogen pro®les are signi®cantly in¯uenced by the
spatial laser intensity distribution, we expect a similar e€ect for the hardness. Fig. 92
shows the surface hardness measured laterally across a homogenized laser spot. A
picture of the spot is also included for comparison. The hardness increases from
about 2.5 GPa for the untreated material to about 6 GPa inside the laser spot. It is
seen that the hardness increase is strictly limited to regions inside the laser spot. A
distinct feature seen here is the decrease of the hardness to values below those of the
untreated material, just at the border of the laser spot. This is explained by anneal-
ing e€ects, where the laser intensity is not high enough to invoke laser nitriding
(probably the temperature just remains below the melting point), but the tempera-
ture rise leads to an annealing and thus softening at the surface.
The surface hardness is increasing with the number of pulses n, although for a
single laser pulse even a softening of the surface is found, only two or more pulses
increase the surface hardness. The hardness given in GPa increases with the number
of pulses n according to M ˆ 2:5…2† ‡ 3:5…3†  …1 exp… n=56…12††) and the max-
imum value reached is about 6 GPa.
The hardness depth pro®les shown in Fig. 93 corroborate the initial increase and
®nal saturation of the hardening e€ect. The hardened depth ®rst increases with the
number of pulses and then saturates at about 1 mm.

Fig. 92. Lateral surface hardness across the laser spot. The optical view of the laser spot is included
(4 J/cm2, 64 pulses).
106 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

7.9. Summary of the homogenized beam irradiation

Not surprisingly the homogeneous spatial distribution of the laser intensity also
involves much more homogeneous treatment results.
Fig. 94 gives a synopsis of the lateral e€ects, the lateral nitrogen, surface and
hardness pro®les across the laser spot. The nitrogen distribution follows the spot
dimensions. The surface pro®le and hardness increase coincide. The softening due to
annealing as well as the piston e€ect make the border inside the laser spot visible,
where the material was molten, approximately at 2 mm, i.e. 0.5 mm inside the edge
of the laser spot.

Fig. 93. Hardness depth pro®les for the homogenized beam irradiation with n pulses (4 J/cm2).

Fig. 94. Comparison of all lateral measurements: surface pro®le, surface hardness and nitrogen con-
centration.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 107

The homogeneous pro®le, causing the absence of intensity ¯uctuations and piston
e€ect, enables the accurate determination of the laser nitriding threshold. Strong
nitriding sets in at 2.4(1) J/cm2, exactly where the simulations predict that the boil-
ing temperature is reached at the surface.

8. Phase formation

First an overview about the available MoÈssbauer data in the Fe±N system will be
given with special emphasis on the MoÈssbauer spectroscopic data of austenite and "
nitride. Then the results of the phase analyses for laser nitrided iron, irradiated with
the raw and with the homogenized beam, are presented, including further informa-
tion which can be obtained from the MoÈssbauer phase analysis [496,497].

8.1. Overview on MoÈssbauer data of the Fe±N system

There exists a large number of MoÈssbauer studies in the Fe±N system. Recent
overviews are given by Schaaf et al. [44,45,47,498]. A collection of the MoÈssbauer
parameters for the corresponding phases of the Fe±N system can be found there as
well as in the original literature [499±504,162]. The recently discovered new FeN
phase and its hyper®ne parameters are still under investigation [148,178,152,153].
Fig. 95 gives an overview of the MoÈssbauer spectra of the most important phases of

Fig. 95. MoÈssbauer overview of the Fe±N system [45].


108 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

the Fe±N system [44,45]. Due to their importance for the laser nitriding process, the
austenite and the e-nitride are described below in greater detail.

8.2. Austenite

The nitrogen austenite, or g-Fe(N), is a well-known phase for MoÈssbauer spec-


troscopy. It is paramagnetic over its whole composition range [161]. The ordering of
nitrogen interstitials as well as the proper ®tting of the hyper®ne parameters are still
in discussion [499,500,505,506,238,507]. It is now established that nitrogen austenite
should be ®tted with one singlet (A0) and one or two doublets (A1 and A2),
depending on the nitrogen concentration [506,44]. In iron±carbon austenite, an
ordering of the carbon interstitials is observed, originating from the repulsive inter-
action of the carbon interstitials [499,508], thus allowing an iron atom having a
maximum of one nearest carbon neighbor (nn) and up to four next nearest carbon
neighbors (nnn). In iron±nitrogen austenite there are no repulsive forces and no
ordering, since N atoms are smaller than C atoms, leading to a random distribution
of nitrogen atoms on the interstitial sublattice (octahedral sites) [505,499]. Although
some authors claim a weak ordering in the nitrogen austenite [506], no clear proof
has been given yet.
If we write the nitrogen austenite as FeNyN , with yN ˆ cN =…1 cN † where cN is the
nitrogen concentration, we have with yN the probability that a certain octahedral
site is occupied by a nitrogen atom. Since an iron atom has 6 nearest neighboring
octahedral interstitial sites, the probability pn …yN † for an iron atom to have n nearest
nitrogen neighbors is given by
 
6
pn … yN † ˆ …yN †n …1 yN †6 n for n ˆ 0; . . . ; 6: …119†
n

Taking the maximum solubility of cN ˆ 10:3 at.%, or y ˆ 0:115, we should have up


to three nearest nitrogen neighbors, while p4 remains as small as 0.1%. The in¯uence
of next-nearest nitrogen patoms  on the hyper®ne parameters should be negligible
because their distance is 3 times larger than that of the nearest neighbors. Conse-
quently, they cannot be resolved in the MoÈssbauer spectra. The hyper®ne interac-
tions not only re¯ect the number but also the arrangement of the neighbors. Taking
into account the di€erent con®gurations of the nitrogen neighbors and performing
Point Charge Model (PCM) calculations [44] for the electric ®eld gradients (EFG) it
has been possible to achieve a proper assignment of the subspectra [506,44]. The
nitrogen content of the austenite can be obtained from the relative fractions of the
MoÈssbauer subspectra with surprisingly high accuracy [44,506] by:
s
6 f …A0†
yN ˆ 1 P ; …120†
i f …Ai†

where f …A0† is the relative area of the single line without any nitrogen neighbor and
f …Ai† are the relative areas for the iron sites having at least one nitrogen neighbor.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 109

8.3. Epsilon-nitride

The " iron nitride has a hexagonal crystal structure (P312 or P6322) [102,133].
Nitrogen is soluble between 15 and 33 at.%. At room temperature, the " phase is
ferromagnetic for nitrogen concentrations of 17.5±30 at.% and paramagnetic for
16.1±17.5 at.% and 30±33.3 at.% [163,160]. Increasing the nitrogen concentration
from 33 to 33.2 at.% leads to an anisotropic distortion of the " nitride lattice, and
the orthorhombic z-Fe2N line phase is formed.
The values of the MoÈssbauer parameters of the "-Fe2 N1 z nitride in the range of
0  z  1=3 are reported in the literature [44,504,178,509±511]. According to Jack
[102,511], an ordering of the nitrogen atoms is e€ective, leading to the following
probabilities of the di€erent iron sites with z from "-Fe2 N1 z :

p…FeIII† ˆ 1 3z …121†

p…FeII† ˆ 3z; …122†

where the sites are named according to the number of nitrogen neighbors, FeIII
having 3 nitrogen neighbors and FeII having 2 nitrogen neighbors. The sites FeI and
Fe0 should not occur in this concentration range, if the ordering is e€ective.
The ordering is displayed in Fig. 96. As early as 1952 Jack proposed [102] the
following `rules' for the occupation of the nitrogen interstitials: (i) in each plane a
maximum of 2 nitrogen atoms are allowed and (ii) an occupied interstitial site A, B,
or C is blocking the same sites A, B or C in the plane immediately above and below.
Fig. 97 shows the variation of the site fractions with the nitrogen content in the
e-nitride. It is easily recognized that the fractions nicely follow the solid lines,

Fig. 96. Schematic drawing of the " lattice with the ordering of the N interstitials (adopted from [28]).
110 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

representing the short range order proposed by Jack [102]. One also notices, that the
measured fractions are far away from the dotted lines for random (binomial) site
occupation. Nevertheless, the FeI site should not occur. Especially the perfect coin-
cidence of the FeIII site implies that MoÈssbauer spectroscopy may be the most
accurate method for determining the composition of "-Fe2 N1 z , provided the spec-
tral areas (no thickness e€ects, identical or known Debye±Waller-factors) are well
known and the ordering is e€ective (no quenched samples). Chen et al. [511] have
doubted that the ordering model is e€ective in the nitride, but if one takes into
account that the authors have not well resolved the paramagnetic spectra, i.e. not
®tted the FeII site, their values still agree with the ordering within the error limit.
The degree of ordering in the " nitride is still being discussed [44,142,141,140].
Although there is now no doubt, that an ordering is e€ective, it is di€erent from the
one proposed by Jack [102] and evidenced by MoÈssbauer investigations [45,140,128].
The solely appearance of the FeII subspectra in Fe3N, for example, clearly proves
the ordering. With the aid of nitrogen-absorption isotherms Somers and coworkers
found a transition of the ordering from the Fe3N to the Fe2N region [141,142].
Jacobs and coworkers prefer the structure type P321 for Fe3N with additional
occupation of nitrogen sites, i.e. a preferential site occupation [140,133]. The MoÈss-
bauer results shown above endorse the suggestion of Somers and coworkers because
there is a clear deviation from the full ordering around z ˆ 0:25, which hints to a
transition in the ordering mechanism. It also became evident that the ordering pro-
posed by Jack fails describing new XRD and neutron di€raction data. MoÈssbauer
spectroscopy usually only measures the short range order and only in very rare
examples also a long range order (e.g., in 0 -Fe4N). Nevertheless, it was shown, that
MoÈssbauer phase analysis can accurately determine the nitrogen content in the

Fig. 97. Ordering in "-nitride: areas of the subspectra (open symbols from gas nitriding, ®lled symbols
from magnetron sputtering, solid lines for ordering model, dashed lines for binomial distribution) [496].
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 111

ordered "-nitride by the ordered model of Jack, even without knowing the exact
ordering in this phase.
Furthermore, the magnetic hyper®ne ®eld is sensitive to the nitrogen content in
the " phase. This is shown in Fig. 98, where only gas-nitrided samples are included,
because they are slowly cooled [140] and thus in thermal equilibrium. The magne-
tron sputtered samples, especially those prepared at room temperature, show
annealing e€ects as published by Niederdrenk and coworkers [178]. The hyper®ne
®elds Bhf (in Tesla) can be well described by a linear equation in the range 0:14 
z  0:32 before TC falls below room temperature for z < 0:12 and z > 0:44 (see Figs. 4
and 98):

Bhf …FeI† ˆ 19:05…21† ‡ z  38:4…3:5†


Bhf …FeII† ˆ 11:71…66† ‡ z  41:1…2:8†
Bhf …FeIII† ˆ 2:48…62† ‡ z  35:7…2:6†: …123†

Thus the magnetic hyper®ne ®elds can also be used to determine the nitrogen con-
centration in the "-nitride. In conclusion we have two possibilities for an accurate
determination of the nitrogen content in the " phase: (i) the relative fraction of the
FeIII site and (ii) the hyper®ne ®eld Bhf of the FeII and FeIII sites. A comparison of
the nitrogen content determined in this way by MoÈssbauer spectroscopy with those
obtained via ion beam analysis (IBA) has been carried out by Niederdrenk et al.
[178]. A very good agreement was found.

8.4. Results for laser nitrided iron

So far, no phase analyses for the laser nitrided iron were presented. For technical
applications it is of great importance which phases are formed in the surface after

Fig. 98. Hyper®ne ®elds Bhf in " nitride (only gas-nitrided samples) [496].
112 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

the laser nitriding process. The phase analyses presented below were performed by
MoÈssbauer spectroscopy, which enables an accurate quantitative phase analysis, as
just reported. Additional analyses were performed by XRD. The information depth
is di€erent and the quantitative accuracy of XRD is more limited. Therefore, mainly
the MoÈssbauer results are presented here [51,52,74].
Fig. 99 compares the CEM and CXM spectra obtained for laser nitrided iron,
using 32 pulses (8  4) of the raw beam at 4 J/cm2. There is a clear di€erence in the
structure of both spectra: The CEM spectrum exhibits a superposition of the sub-
spectra for the , 0 , and " phases. The CXM spectrum only needs the , 0 and
subspectra to be ®tted. The CEMS analysis gives 20(3)% ferrite/martensite, 22(2)%
austenite and 58(3)% "-nitride, while the CXMS result is 89(2)% ferrite/martensite
and 11(2)% austenite. Taking into account the information depth for both mea-
surements (about 150 nm for CEMS and 10 mm for CXMS) one can conclude that
the " phase is limited to the surface region, followed by the phase, present also in
deeper regions. The depth of the modi®ed layer can be estimated to be about 1.4(6)
mm, in agreement with the values determined by the simulation and the brittle frac-
tures (e.g., Fig. 39).
Fig. 100 shows three CEM spectra obtained for di€erent laser ¯uences H and the
strong in¯uence of H on the phases formed. The " phase is not in thermodynamic
equilibrium as observed on the high amount of its Fe-I subspectrum. Nevertheless,
the hyper®ne ®elds can be used to estimate the nitrogen content of this phase (see
Eq. (123)), which ranges from 20 to 30 at.%, covering almost the complete range of
existence. Sometimes also the paramagnetic doublet is found, indicating that part of
the e phase has more than 30 at.% nitrogen. The nitrogen concentration of the
austenite can reach 12(1) at.%, which is higher than the maximum solubility in the
equilibrium phase diagram. The nitrogen content in the austenite is highest, just
before the " phase starts to form, and decreases as more and more " is formed. For

Fig. 99. Comparison of the CEM and CXM spectra for laser nitrided iron (4 J/cm2, raw beam, 32 pulses
(8  4), spot size 3.8  4.7 mm2): (a) CEMS, (b) CXMS.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 113

2 J/cm2 only small changes relative to the unirradiated a-Fe are observed. For
higher energy densities, the original ferrite fraction has almost disappeared. The
changes of the fractions are illustrated in the summary of the CEMS phase analyses
shown in Fig. 101.
The phase composition in the surface regions of the laser nitrided samples starts to
change at the threshold laser ¯uence of about 2 J/cm2 and soon reaches saturation.
The phase composition hardly changes for laser ¯uences of 4 J/cm2 or more, in
agreement with the simulations presented in Section 4.3. The fraction of the " phase
reaches about 55%.
The phase composition in the laser nitrided samples is very sensitive to the spot
size and the homogeneity of the spatial laser intensity distribution as seen by com-
paring the CEM spectra in Figs. 99(a) and 100(b). Although all treatment para-
meters except the laser spot size are identical, the phase composition is di€erent. For
the smaller spot size the abundance of the " phase is about 58(3)%, whereas for the

Fig. 100. CEM spectra of laser nitrided iron for various ¯uences H (raw beam, 32 pulses (8  4), spot size
4.6  5.6 mm2).
114 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

larger spot size it is reduced to 43(3)%. As the piston e€ect and the amount of fall-
out are increasing for smaller spot sizes, as seen in Section 6.7, the amount of the "
phase increases for the meandering treatments presented here because the fall-out
consists of the " phase as shown in Section 6. This assumption may be proven by a
laser nitriding experiment performed with the homogenized beam. Due to the
absence of fall-out, no " phase should appear there.
Fig. 102 compares the CEM spectra obtained after the laser nitriding of iron with
the raw and the homogenized beam. It was shown that the homogenized beam leads

Fig. 101. Phase fractions in laser nitrided iron as function of the energy density H (raw beam, 32 pulses
(8  4)).

Fig. 102. Comparison of the CEM spectra obtained for 32 (8  4) pulses at 4 J/cm2: (a) raw beam;
(b) homogenized beam.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 115

to the formation of the a00 phase (15(3)%), while the raw beam irradiation produces
the " phase (44(3)%) instead. The " phase is not fully ordered with respect to the
nitrogen interstitials, as indicated by the high fraction of the Fe-I site. The nitrogen
concentration of the " phase is about 28(2) at.%, as obtained from the hyper®ne
®eld Bhf …FeII† via Eq. (123). The g-Fe phase has a fraction of 37(2)% and contains
9.6(6) at.% nitrogen for the homogenized beam. The raw beam exhibits 31(3)%
with a nitrogen content of 7.6(7) at.%. A rough estimate of the mean nitrogen con-
centration in the surface based on the MoÈssbauer phase analysis gives 6 at.% for the
homogenized beam and 15 at.% for its raw beam counterpart. The formation of the
" phase is attributed to the deposition of fall-out outside the laser spot. MoÈssbauer
and RNRA measurements proved that this fall-out consists of " phase with a
nitrogen concentration of about 30 at.%. The absence of the " phase for samples
irradiated with the homogenized beam may not indicate that it has not been formed
at all. The fall-out, which is deposited inside the laser spot, is dissolved and inter-
mixed in the liquid layer, at least with the subsequent pulse. This means that the
thickness of "-fall-out is very small, too small to be visible in the MoÈssbauer mea-
surements. The fall-out deposited outside the laser spot for the raw beam treatment
remains there and even accumulates during the meandering laser treatment. Its
phase fraction of about 50% would then correspond to an " layer of about 32 nm
thickness.9
Further experiments for the phase formation were carried out with the 248 nm, 30
ns, KrF excimer laser (see Section 5.2). Here, the used spot size (2  2 mm2) was
much smaller since the maximum pulse energy was only 0.3 J/cm2. The laser ¯uence
was adjusted with a variable attenuator and a beam homogenizing unit was used for
these experiments. The energy densities of 2.1, 2.5, and 3.1 were employed using up
to 100 pulses for the tile-like irradiations (1  1) [52]. Only a weak nitriding for this
high number of pulses, was observed at the laser ¯uence of 2.1 J/cm2, i.e. the same
threshold energy density as reported before, was obtained.
Fig. 103 displays the CEM spectra obtained for 5, 12 and 32 pulses at 3.1 J/cm2.
The spectra are again composed of , , 0 and ". The amount of " is increasing with
the number of pulses. This is visualized in Fig. 104, where the phase fractions as
obtained from the CEM spectra are shown. The 00 is limited to low number of
pulses, for higher pulses it disappears to the expenses of the " phase. This would
suggest the transformation of " with a low nitrogen content into 00 , as suggested by
Rochegude and Foct [137].
The signi®cant di€erence to the previous homogenized beam irradiation is the
high amount of " formed here. This might be caused by the much smaller spot size,
also leading to a remarkable piston e€ect and fall-out generation, although the beam
is homogenized. The analogue can be observed for the phase formation with the
laser energy density. The highest amount of 00 is 33(3)%, 41(2)% for the , and
69(3)% for the " phase.
These samples were also analyzed by RNRA, so the mean nitrogen content in the
topmost 150 nm obtained from RNRA and the phase fractions obtained by CEMS

9
With the assumption that this layer is composed of pure " only.
116 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Fig. 103. CEM spectra of laser nitrided iron at 3.1 J/cm2 (248 nm, 33 ns).

Fig. 104. Phase fractions against number of pulses (248 nm, 33 ns, 3.1 J/cm2).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 117

can be compared. This comparison is given in Fig. 105. The phases are formed
according to the phase diagram, although their occurrence is shifted towards lower
concentrations. The 00 already appears around 6 at.%, despite its usual composition
of 11.1 at.%. The fraction of already starts to decrease and " starts to form before
the maximum solubility value of 10.3 at.% is reached.
Again, the phase formation is most in¯uenced by the energy density and the
spatial homogeneity of the laser beam, as the latter determines the amount of
fall-out generated. The fall-out is most probably responsible for the formation of a
high amount of the " phase. Without this fall-out deposition outside the laser spot,
no " phase can be detected in the laser nitrided samples.

9. In¯uence of the nitrogen gas pressure

This chapter describes ®rst experiments carried out in order to resolve the in¯u-
ence of the ambient nitrogen gas pressure pN for the laser nitriding e€ect [47,75,56],
i.e. for the surface pro®les, the nitrogen take-up and the phase formation. These
experiments were carried out under the premises that higher gas pressures would
enhance the nitriding e€ect. Rykalin [512] found an exponential decrease of the laser
transmission to the target with increasing pressure. However with the increase in gas
density, also the plasma pressure acting at the surface should increase [332], leading
to enhanced nitriding. Furthermore, Bettis [292] reported the weak lowering of the
threshold intensity for breakdown with the pressure Ithr / pN2=3 .

9.1. Surface morphology

The ®rst expectation was that the lateral material transport by the piston
mechanism is in¯uenced by the gas pressure. The homogenized beam did not produce

Fig. 105. Comparison of phase fraction and nitrogen content (248 nm, 33 ns).
118 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

a measurable piston e€ect for the 32 pulses used here. Therefore, its pressure
dependence could not be determined. Fig. 106 shows the surface pro®le obtained for
the raw beam irradiation at 4 J/cm2 with 32 pulses (spot size 6.4  5.4 mm2). The
piston e€ect is less pronounced as compared to the irradiation with smaller spot
sizes (e.g., Fig. 37), but nevertheless, the in¯uence of the gas pressure on the piston
e€ect is clearly seen. The mean lowering of the surface in the center of the laser spot,
z is calculated from these pro®les. z starts from about 1.2 mm for the low
pressure region, then decreases with increasing gas pressure and vanishes at around
0.7 MPa. At the higher pressure of 1 MPa the piston e€ect increases again to z ˆ
0:4…1† mm. An increased absorption of the laser irradiation in the plasma may be
one of the reasons for the reduced piston e€ect. This would also explain the decrease
of the lateral extension in the surface pro®les with increased pressure. Between 0.7
and 1 MPa there might be a transition in the plasma behavior (LSC to LSD). The
following results also support the assumption of this transition to happen in this
pressure region.
When the laser treatment is carried out by meandering the laser spot across the
whole sample the piston e€ect may lead to periodic structures as demonstrated
in Fig. 107. The observed periodicity along the y-axis is stronger than that along the

Fig. 106. Pressure dependent surface pro®les of laser nitrided iron (4 J/cm2, raw beam, 32 pulses, single
spot, spot size 5.4  6.5 mm2). The pro®les are shifted.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 119

x-axis (compare also Fig. 22). From the distances of the peaks10 in both directions
we exactly ®nd the value for the shift settings y ˆ 0:50…2† mm and x ˆ 1:31…11†.
These values agree well with the settings for the meandering treatment y ˆ 4:2
mm/8 = 0.52 mm and x ˆ 5:4 mm/4 = 1.35 mm.
As a consequence, this periodicity with the high surface corrugations also leads to
a high surface roughness Ra as displayed in Fig. 108. Ra ®rst increases fast with the
gas pressure, followed by an only slow increase over the region from 0.1 to 0.7 MPa.
Again, above the latter value there is a drop in the surface roughness. The change of
the roughness for the homogenized single spot irradiation is independent of the gas
pressure.

9.2. Nitrogen concentrations

The nitrogen concentrations measured by RNRA after the laser nitriding also
depend on the nitrogen gas pressure during the laser treatment. Fig. 109 shows the
lateral nitrogen distributions across a homogenized laser spot irradiated at various
nitrogen gas pressures. The height of the concentration plateau in the center of the
spot is hardly changing with the gas pressure. Only for the lowest pressure of 0.05
MPa and the highest pressure of 1 MPa this value is lower than for the others.
Furthermore, the treatment at the highest pressure of 1 MPa also led to a narrower
lateral nitrogen distribution. The pro®le becomes smaller and sharper. This again
indicates a transition in the laser plasma behavior between 0.7 and 1 MPa nitrogen
pressure.

Fig. 107. Surface pro®les for the meandering treatment of iron: (a) along the y-axis, (b) along the x-axis.
(4 J/cm2, raw beam, 32 pulses (8  4), 1.0 MPa N2) (spot size a  b ˆ 4:2  5:4 mm2).

10
Fourier transformation for scan (a) gives a well de®ned damped sine, its frequency is the mean peak
distance.
120 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

The corresponding nitrogen depth pro®les measured in the center of the spots are
given in Fig. 110. They all show high surface concentrations which rapidly fall o€ at
about 20 nm and then show the normal di€usion pro®le. There, deeper in the sample
the nitrogen concentration raises with the pressure up to 0.7 MPa, but then decreases
again for 1 MPa.
This behavior is also found in the mean nitrogen content hcN i of the topmost 300
nm as a measure of the total nitrogen take-up, which is shown in Fig. 111 versus pN .
The tendency mentioned above is found again and the take-up of nitrogen increases

Fig. 108. Roughness as a function of the nitrogen pressure (4 J/cm2, raw beam, 32 pulses (8  4), stainless
steel).

Fig. 109. Pressure dependent lateral nitrogen pro®les for homogenized irradiation. (4 J/cm2, homogenized
beam, 32 pulses, single spot).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 121

from 0 at.% for 0 MPa to 20 at.% for atmospheric pressure and then only slowly
increases with the nitrogen pressure up to 0.7 MPa and decreases for 0.1 MPa again.
Fig. 112 summarizes the parameters c2 and z2 obtained from the ®ts of the nitro-
gen depth pro®les as shown in Fig. 110 according to the di€usion pro®le de®ned in
Eq. (110). The parameters c0 ˆ 5…1† at.% and fall out thickness z1 ˆ 8…3† nm almost
do not change with pressure. c2 jumps steeply at 0.05 MPa and then slowly increases,
but is lowered again for 1 MPa. The di€usion length z2 shows a logarithmic beha-
vior (excluding the 1 MPa point) according to z2 ˆ 304…16† nm  log…pN =15…2† hPa†,
thus the nitriding has a threshold of 15(2) hPa.

Fig. 110. Nitrogen depth pro®les for various nitrogen gas pressures. (4 J/cm2, homogenized beam, 32
pulses, single spot).

Fig. 111. Mean nitrogen content in the topmost 300 nm as obtained from the depth pro®les (Fig. 110).
122 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

The behavior of c2 is not surprising. As long as LSC plasma waves are generated,
1=3
the plasma pressure hardly changes with the gas pressure (pp / 1=3 gas / pN ) and
p
consequently the nitrogen solubility c2 (c2 / pp ) increases weakly with the gas
pressure. The gas pressure modi®es the dynamics of the plasma expansion and thus
the time of the plasma-surface interaction leading to changes of di€usion times and
length z2 . A theoretical description of these plasma dynamics and the logarithmic
behavior of z2 needs to be done. The systematic drop found between 0.7 and 1 MPa
may be caused by the transition from the LSC wave to a LSD wave, due to the
increased density and thus increased laser absorption in the plasma [332,320]. The
single spot irradiations carried out with the raw beam (spot size 4.5  5.5 mm2)
show similar results and dependencies as the homogenized beam irradiations
presented here.
Fig. 113 illustrates plasma images for homogenized and the raw beam irradiations
at the gas pressures of 0.1 and 0.9 MPa. The light emission from the laser produced
plasma above the iron surface gives information about the spatial extension of the
plasma. The greyscale is proportional to the light intensity, but unfortunately the
dynamic range of the camera was very low so that not all details can be resolved.
The plasma ignition starts earlier in time for the raw beam irradiation. The lateral
extension of the plasma for the raw beam is larger in both directions, than for the
homogenized beam irradiation. The plasma luminosity is clearly visible for about
300 ns, before it starts to decline, for both cases. Maybe the plasma last some tens of
nanoseconds longer for the raw beam irradiation, but the plasma is better con®ned
to the surface for the homogenized beam.
Concerning the gas pressure pN , the higher pressure also leads to a stronger con-
®nement of the plasma at the surface and the lateral expansion is slowed down (in z-
and x-direction). Thus, it can be expected that the time when the high pressures are
acting at the liquid surface is increased, which explains the increased value of z2 , as
shown above.

Fig. 112. Pressure dependence of the di€usion parameters for the nitrogen depth pro®les.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 123

It remains to investigate if the meandering treatment produces di€erent results due


to the overlapping piston e€ect and lateral material transport. Phase analyses via
MoÈssbauer spectroscopy and XRD were carried out for iron samples treated at
various gas pressures using the meandering technique. Fig. 114 displays the nitrogen
depth pro®les obtained by RNRA for laser nitrided iron using the raw beam at 4 J/
cm2 and 32 pulses by meandering (8  4) the whole surface. The nitrogen depth
pro®les are very sensitive to the gas pressure. Again, they are steep at low pressures
and ¯atten for increasing pressure. All surface concentrations remain very high at
values above 30 at.%, only at 1 MPa the surface concentration is lower at a value of
11 at.%. The concentrations in greater depths ®rst increase with the pressure until a
maximum is reached at 0.1 MPa and then decrease again.
This observation is summarized in Fig. 115 for the three cases. The mean nitrogen
concentration in the topmost 200 nm, shown there, is sharply increasing with the
pressure to reach a maximum of 18 at.% at atmospheric pressure and then decreases

Fig. 113. Plasma images taken for the homogenized beam and raw beam irradiations at 0.1 and 0.9 MPa.
The time after pulse start is given in (ns) (4 J/cm2, lateral spot size: 5 mm for homogenized beam and 5.4
mm for raw beam).
124 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

again upon further increase in the gas pressure, for the raw beam meandering
treatment. The threshold pressure of 10(9) hPa is obtained when linearly extra-
polating the logarithmic pressure scale to a vanishing nitrogen content as reported
above for the homogenized irradiation. The maximum nitrogen take-up is almost
identical …hcN i ˆ 17…1† at.%) for all treatments, but there is a di€erence in the
nitrogen pressure at which the maximum nitrogen take-up occurs for the meandered
samples as compared to the single spots. For the meandered samples, the maximum
is located at atmospheric pressure, whereas it is 0.7 MPa for the single-spot irradia-
tion. Most probably, this is due to the laser generated fall-out, which is re-dissolved

Fig. 114. Pressure dependent nitrogen depth pro®les for iron (4 J/cm2, raw beam, 32 pulses (8  4)).

Fig. 115. Comparison of the mean nitrogen concentration in laser nitrided iron (4 J/cm2, raw and
homogenized beam, 32 pulses, single and meander).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 125

into the surface by subsequent pulses. The generation of fall-out is very sensitive to
the nitrogen gas pressure.

9.3. Phase formation

The results from the RNRA measurement now can be compared with the phases
produced at di€erent pressures.
The CEM spectra of four samples nitrided at 0.025, 0.1, 0.7, and 1 MPa are shown
in Fig. 116. All spectra are ®tted by a combination of the subspectra for the , 0 , ,
and " phases. The spectrum taken for 1 MPa also contains the three subspectra for
the 00 phase. The phase fractions extracted from the MoÈssbauer phase analysis are
summarized in Fig. 117. The pressure dependence for the fraction of the phase
looks quite similar to the mean nitrogen take-up shown in Fig. 115. The " phase has
its maximum fraction of about 30% for pN ˆ 0:1 MPa, and then linearly decreases
with the pressure and vanishes at 1 MPa, where an almost 20% fraction of the 00
can be found.
Having now the phase analysis and the nitrogen content, these two results can
again be combined into the diagram shown in Fig. 118. The nitrogen content was
determined from the RNRA results by integrating over the topmost 150 nm, the
mean information depth of the CEMS measurement. The phase has its maximum
fraction of 10.5 at.%, which according to the phase diagram, is the value of its
maximum solubility. The 00 phase occurs already about 2 at.% below its nominal
composition of 11.1 at.%. The " phase starts, when the nitrogen concentration
exceeds the maximum solubility of the phase. The 00 phase seems again to be

Fig. 116. CEM spectra of laser nitrided iron: various nitrogen gas pressures (4 J/cm2, 32 pulses (8  4),
raw beam).
126 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

formed when there is " with very low nitrogen concentrations which then is trans-
formed into the 00 phase [137].

9.4. Microhardness

Finally we discuss the hardness obtained by the laser nitriding as a function of the
nitrogen gas pressure. Fig. 119 summarizes the hardness depth pro®les. As expected
from the results shown so far, the maximum hardness of more than 10 GPa (com-

Fig. 117. Phase fractions as function of the nitrogen gas pressure (4 J/cm2, 32 pulses (8  4), raw beam).

Fig. 118. Comparison of the phase fractions with the nitrogen content as determined by RNRA (4 J/cm2,
32 pulses, meander 8  4, raw beam).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 127

parable to more than 1000 HV 0.1) is obtained for atmospheric pressure. For low
pressures and for 1 MPa the hardness increase is low and the hardening depth is
small. For 0.1, 0.4 and 0.7 MPa the hardening depth reaches almost 2 mm.
Looking at the technically important surface hardness, one recognizes in Fig. 120,
that the surface hardness Ð determined here by indentation up to about 150 nm Ð
exactly follows the mean nitrogen content in the surface. The surface hardness
reaches values of almost 6.5 GPa. Thus, the meandering treatments produce a
higher surface hardness than the single spot homogenized beam irradiations, but the
same as the single spot raw beam irradiations. Also the development of the surface

Fig. 119. Hardness depth pro®les of iron laser nitrided in various nitrogen gas pressures (4 J/cm2, raw
beam, 32 pulses (8  4)).

Fig. 120. Pressure dependence of the surface hardness (4 J/cm2, raw beam, 32 pulses (8  4)). The mean
nitrogen content is shown for comparison.
128 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

roughness shows the same behavior with the gas pressure. The threshold pressure
obtained from these measurements is 15(10) hPa, the same threshold as obtained
before.

9.5. Summary of pressure dependence

Summarizing the pressure dependent experiments, we conclude that the gas pres-
sure mainly in¯uences the plasma dynamics. If the pressure is too low (pN < 0:1
MPa) the plasma waves cannot fully develop and above 0.7 MPa there is probably a
transition from the LSC to the LSD regime. The importance of the piston e€ect is
decreasing with the gas pressure, which might be caused by the more dense and more
homogeneous plasma acting at the liquid surface. The time of intense plasma-
surface interaction increases with the gas pressure. Before drawing ®nal conclusion
from the pressure dependence, more experiments are currently being performed and
a detailed treatment of the plasma wave dynamics would be necessary.

10. In¯uence of alloying elements

Usually, no tools or functional parts are made of pure iron. Mainly steel or cast
iron is used as a base material. Therefore, when looking for technical applicability,
also the in¯uence of alloying elements on the laser nitriding process is of great
interest. Besides the pure iron various carbon steels and an austenitic stainless steel
were used for the laser nitriding experiments in order to investigate the in¯uence of
the material itself.

10.1. In¯uence of the carbon content

It has been shown in the previous chapters [44,45,47] that the laser nitriding pro-
cess in pure iron leads to the formation of the phases " and and sometimes of some
00 . The in¯uence of the carbon content was examined by investigating various plain
carbon steels with di€erent carbon contents ranging from 0 to 1.05 wt.% carbon
(Fe, Ck15, Ck45, C60, C80, C105). We describe here the results for the laser nitrid-
ing process, i.e. phase formation and nitrogen take-up as a function of the carbon
content. All treatments were carried out with the 4 J/cm2 raw laser beam at atmo-
spheric pressures. The whole surface of the samples was covered by meandering the
laser spot (8  4). A treatment in pure Ar atmosphere was included, in order to
di€erentiate the `normal' fast heating and quenching e€ects from the nitriding
e€ects.
Fig. 121 shows the XRD pattern of steel C80 (containing about 0.8 wt.% carbon)
before and after laser irradiation in argon and nitrogen atmosphere. The untreated
C80 sample (a) exhibits only the peaks of the original a-Fe. The irradiation in Ar led
to a signi®cant broadening of the peaks and to the appearance of some small peaks
attributed to the formation of a small amount of g-Fe, the retained austenite from
the hardening process. The broadening is due to the martensite formation and shock
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 129

hardening [513,514] accompanied by stress formation in the treated surface. The


sample irradiated in nitrogen atmosphere exhibits more dramatic changes. Most
obvious is the huge amount of austenite present in the treated layer. There is also a
clear indication for the presence of "-nitride.
This can now be compared with the results of the MoÈssbauer phase analysis.
Fig. 122 displays the CEM spectra obtained for the C80 in argon and nitrogen
atmosphere. The results of the XRD pattern are qualitatively also found in the
CEM spectra. The sample irradiated in argon shows a small amount of the austenite
singlet and doublet in the center in addition to the , 0 sextets. The irradiation in
nitrogen led to a strong increase of the austenite and especially the " subspectra. One
should remember that the information depth for CEMS is about 150 nm, whereas
for XRD it is about 1 mm when using a ®xed incidence angle of 5. This explains why
CEMS enhances the " phase, which is limited to the surface region. The CEMS and
CXMS measurements of the steel C80 are displayed in Fig. 123. The fraction of the
latter is larger than for pure iron treated under the same conditions. This will be
discussed in detail along with the quantitative phase analyses given below.
The depth of the laser nitrided layer can be estimated, when comparing the con-
version electron and the conversion X-ray MoÈssbauer spectra. In Fig. 123 this is

Fig. 121. XRD spectra of steel C80: (a) untreated; (b)laser treated in Ar; (c) laser treated in N2 (32 pulses
(8  4), 4, raw beam) (XRD with ®xed incidence angle of 5 ).
130 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Fig. 122. Comparison of laser treated C80: (a) untreated; (b) in argon; (c) in nitrogen.

Fig. 123. Comparison of CEM (a) and CXM (b) spectra of laser nitrided C80 (4 J/cm2, raw beam, 32
pulses (8  4), 0.1 MPa).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 131

done for the laser nitrided C80. While the CEM spectrum shows the presence of a
high amount of the " and g-Fe(N), the CXM spectrum with its larger information
depth shows only a small amount of g-Fe(N). Thus, it can be concluded that the
presence of the " is limited to the very surface. From the absorption coecients a
transformed layer in the order of about 1 mm can be deduced.
The di€erent phase formation for irradiation in argon and nitrogen re-appears in
the hardness depth pro®les, as shown in Fig. 124. The highest hardness with a
maximum of 7.3 GPa is achieved for the irradiation in argon because the martensite
can be much harder than the and "-phases. The hardening depth Ð here deter-
mined as the depth where the hardness reaches the hardness of the untreated
material Ð is about 3 mm. The sample irradiated in nitrogen has only a maximum
hardness of 5.7 GPa, but the hardening depth of 4 mm is somewhat larger.
Coming back to the question, how the carbon content in¯uences the laser nitrid-
ing process, Fig. 125 shows CEM spectra obtained for laser nitrided Fe, Ck15, C80
and C105. It immediately becomes obvious that the carbon content of the samples
strongly in¯uences the MoÈssbauer spectra.
It could be expected that the transition from the binary Fe±N system to the tern-
ary Fe±N±C system may cause some diculties in the MoÈssbauer phase analysis.
Nevertheless, all spectra are well ®tted with the same set of subspectra for the phases
, , " and 00 of the Fe±N system (see Chapter 2) with varying contributions. The
used MoÈssbauer
 parameters can be found in Refs. [44]. The carbonitride "-
Fex N1 y Cy , which probably evolves here, is hardly systematically investigated, but
®rst results show that the appearance of this phase in the MoÈssbauer spectra is more
in¯uenced by x than by y [497]. The additional carbon should not signi®cantly
change the parameters for the austenite. A possible presence of cementite () cannot
be excluded due to the strong overlap with the " subspectra [515], but as thermo-
dynamics favors the formation of the " phase [114] we assume that no cementite is

Fig. 124. Comparison of the hardness depth pro®les for laser treated C80.
132 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

present in the CEM spectra. Furthermore, the additional carbon stabilizes the and
" phase, but decreases the existence range of the and 0 phases [114].
The resulting MoÈssbauer phase analyses are summarized in Fig. 126. An increase
of the " phase fraction with the carbon content can be observed, from about 40%
for pure Fe to about 60% for C80 and C105. The fraction of " increases at the
expense of the fraction. The amount of remains almost constant, as well as that

Fig. 125. CEM spectra of laser nitrided samples: (a) Fe; (b) Ck15; (c) C80; (d) C105 (4 J/cm2 (8  4) in 0.1
MPa nitrogen gas).

Fig. 126. Phase fractions in the surface of laser nitrided steel as obtained by CEMS in dependence of the
carbon content.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 133

of 00 . A small amount of oxide (wuÈstite) can be found in some cases. All the CXMS
spectra (not shown) with their larger information depth (10±20 mm) only exhibit the
phases and , the latter rising linearly from about 5% for Fe to 20% for C105.
This is a clear indication that the heat a€ected, i.e. austenized layer, also increases
with the carbon content, due to a lowered austenizing temperature A3, which has a
minimum at the eutectoid point in the Fe±C system at 0.80 wt.% C or to a lowering
of the liquidus temperature with increasing carbon content [131]. This behaviour
should be re¯ected also in the hardness of the samples.
The surface hardness shown in Fig. 127, does not systematically vary with the
carbon content. For the nitrided sample the surface hardness is about 5.5 GPa and
for the samples irradiated in argon it is about 6 GPa, slightly higher.
The hardening depth obtained from the hardness depth pro®les is displayed in
Fig. 128. For the laser nitrided carbon steels the hardening depth increases from
about 1 mm for Fe and Ck15 to 2 mm for Ck45 and to about 4 mm for C80. For C105
the hardening depth is again reduced to 2 mm.
The nitrogen take-up was measured by RNRA and the corresponding nitrogen
depth pro®les for the carbon steels do not exhibit large e€ects of the carbon content.
The surface concentrations are between 20 and 25 at.%, becoming slightly higher
and the gradient steeper with increasing carbon content. The total nitrogen take-up
as the mean concentration averaged over the RNRA nitrogen depth pro®les changes
very little with the carbon content (12 at.% mean concentration in the ®rst 300 nm),
as displayed in Fig. 129. All samples show mean concentrations around 12 at.%.
There is only a slightly higher value for steel C80 (13 at.%). The nitriding mechan-
ism therefore does not seem to play any role. Thus, the increase of the " phase with
the carbon content observed by MoÈssbauer spectroscopy is a result of the thermo-
dynamics being changed by the carbon [114], but not of a changed nitrogen take-up
during the laser nitriding treatment.

Fig. 127. Surface hardness for the various carbon steels irradiated in Ar and N2.
134 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

When looking for the bene®ts of the laser nitriding process, it has to compete with
other processes, such as gas nitriding. For both processes, gas nitriding and laser
nitriding, MoÈssbauer phase analysis shows the formation of the " phase. The di€er-
ence is the additional formation of the phase during the laser nitriding, whereas
the gas nitrided samples show the 0 phase instead. This di€erence easily follows
from the treatment temperatures: 823 K for gas nitriding and up to 4500 K for laser
nitriding [50]. The occurrence of the 00 phase in the laser nitrided samples, however,
is not easy to understand. The fast quenching process should hinder the formation
of the complicated 00 structure [101]. Maybe the transformation of low nitrogen "
into 00 [137] or the annealing of the nitrogen martensite [101] during the treatment
of neighboring areas can explain this.

Fig. 128. Hardening depth for laser nitrided carbon steel.

Fig. 129. Mean nitrogen content in the ®rst 300 nm as a measure for the nitrogen take-up.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 135

Although there are great di€erences in treatment temperature, kinetics and phase
formation, the resulting surface hardness is comparable for both processes. Fig. 130
compares the microhardness depth pro®les of steel C15 after laser nitriding and after
gas nitriding. For both cases a surface hardness of about 5 GPa (HU 1) is obtained.
Nevertheless, the shape of the pro®les is completely di€erent. For laser nitriding,
high hardness values are already obtained at the very surface, which then soon drop
to the value of the untreated material at about 1 mm. The gas nitrided sample is quite
soft at the very surface and then the hardness increases to the maximum at about 0.5
mm. The hardening depth here is in the order of 3±4 mm. In both cases the " phase
should be responsible for the high hardness (M…"† ˆ 7:1 GPa, [136]).
The in¯uence of the carbon content on laser nitriding of iron was investigated.
Laser nitriding leads independent of the carbon content within seconds to nitride
formation, resulting in surface layers of high hardness. The " phase occurs for all
steels and its amount is increasing with the carbon content of the laser nitrided steel.
The achieved hardness values are comparable with the gas nitriding process,
although the hardening depth is smaller. For laser nitriding the hardening depth is
much larger for C80 as compared to the other steels. Maybe the eutectoid compo-
sition with the minimum in the A3 temperature has this bene®cial e€ect.
Carbon has no signi®cant e€ect on the nitrogen take-up during the laser nitriding
process, but the formation of the "-phase (carbonitride) is enhanced. The hardening
depth is also increased for higher carbon contents but is still much lower than for the
gas nitriding process.

10.2. Laser nitriding of stainless steel

All the previous laser nitriding experiments were carried out on iron and steels
with bcc structure. Since it is known that the nitriding of high alloy stainless steel is

Fig. 130. Microhardness depth pro®les obtained by laser nitriding (open squares) and gas nitriding (®lled
circles) of C15.
136 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

extremely dicult [217,211,516] experiments were carried out to investigate the laser
nitriding along this line. The austenitic stainless steel 1.4401 (X5CrNiMo18.10) was
chosen because it is a high alloy austenitic stainless steel with fcc structure.
That the laser nitriding is e€ective also for the stainless steel is easily seen in
Fig. 131, where the nitrogen depth pro®les as revealed by RNRA are displayed for
several nitrogen gas pressures. The depth pro®les have similar shapes as compared
to those of pure iron or carbon steels. The maximum nitrogen concentrations also
reach about 30 at.% at the surface. Again, the nitriding eciency is increasing with
the nitrogen gas pressure until it drops for 1 MPa.
In Fig. 132 the total nitrogen take-up (300 nm) of the stainless steel is shown as a
function of the nitrogen gas pressure pN . A steep rise at about 0.05 MPa is seen,
followed by a slow increase up to 0.7 MPa. A further increase of pN to 1 MPa lowers
the nitriding eciency, most probably due to a transition from a LSC wave to a
LSD wave. The formation of a LSD wave blocks all laser radiation from reaching
the surface, no further nitrogen take-up or di€usion are possible. The nitriding
threshold of the laser ¯uence H and the total nitrogen take-up for laser nitriding of
stainless steel are the same as for iron.
Despite these similarities, di€erences in the phase formation should be expected.
The high alloy austenitic steel will not show the martensitic hardening e€ect, and the
high Cr and Mo content may lead to the formation of chromium nitrides (CrN,
Cr2N), since the anity to nitrogen is much higher for chromium that for iron.
Fig. 133 shows the CEM spectrum of the laser nitrided stainless steel. The laser
treatment led to the appearance of a singlet (S2) and doublet (D) in addition to the
singlet (S1) of the starting material. The hyper®ne parameters are for S1: 1 ˆ
0:11…2† mm/s, S2: 2 ˆ 0:03…2† mm/s, and D: D ˆ 0:04…2† mm/s and "D ˆ 0:44…2†
mm/s. There seems to be a relation between the isomer shift of S2 and D with the
nitrogen content, but further experiments are necessary to test this dependence. The
singlet S1 and the doublet D are attributed to the nitrogen austenite ( -Fe(N))

Fig. 131. Nitrogen depth pro®les for laser nitrided stainless steel (4 J/cm2, 32 pulses (8  4), raw beam).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 137

formed by the laser nitriding behavior. Its fraction clearly follows the total nitrogen
content in the sample as measured by RNRA. The presence of martensite can be
excluded and there are no indications for the presence of " nitride or mixed nitrides.
Pure chromium nitrides are, of course, invisible in the MoÈssbauer spectra, but its
presence was excluded by XRD measurements. Jervis [517] treated also a AISI304
stainless steel with an excimer laser in air. In the MoÈssbauer spectra he also found a
second singlet besides the original in the CEM spectra.
A similar observation is made in the XRD pattern of the laser nitrided stainless
steel as shown in Fig. 134. After the laser nitriding an additional fcc phase with

Fig. 132. Mean nitrogen concentration as function of the nitrogen gas pressure (4 J/cm2, 32 pulses (8 
4), raw beam).

Fig. 133. CEM spectrum of laser nitrided stainless steel (4 J/cm2, 32 pulses (8  4), raw beam, 0.1 MPa).
138 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

shifted lattice parameter is observed, apart from the original fcc phase of the stain-
less steel. There are no indications for other phases, e.g. chromium nitrides. The shift
in the lattice parameter is related to the nitrogen content in the austenite (see Eq.
(4)), shown in Fig. 135. The nitrogen content derived from XRD, behaves the same
way as the nitrogen take-up determined by RNRA and the phase fraction obtained
from CEMS. As there is no martensitic hardening of the stainless steel, one would
expect a smaller increase of the hardness than for iron or carbon steel. Fig. 136

Fig. 134. XRD patterns for laser nitrided stainless steel (4 J/cm2, 32 pulses (8  4), raw beam, 0.1 MPa,
meandering). The insert shows the details of the 220 peak.

Fig. 135. Lattice constant and nitrogen content in the austenite of the laser nitrided stainless steel.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 139

shows the hardness depth pro®les of laser nitrided stainless steel. They look similar
as for the iron and carbon steel samples and also the maximum hardness of about
5.2 GPa is quite close to their values and to that of g-Fe(N) [518,519]. We conclude
that austenitic stainless steel experiences an increase in the hardness, probably by the
combination of stresses induced by the nitrogen incorporation and a shock hard-
ening leading to dislocations and also stresses [513,514,520±522]. Finally, the surface
roughness Ra increases like for the other alloys with the pulse number, as shown in
Fig. 137 for the laser treatment at 4 J/cm2 and 0.1 MPa. Ra increases from the value
of 10(5) nm for the untreated material top250  nm after irradiation with 64 pulses,
according to the function Ra ˆ R0a ‡ R0a  n, where R0a ˆ 280…30† nm.

Fig. 136. Hardness depth pro®les of laser nitrided stainless steel (4 J/cm2, 32 pulses (8  4), raw beam,
0.1 MPa).

Fig. 137. Surface roughness of laser nitrided stainless steel (4 J/cm2, meander, 0.1 MPa).
140 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

But there are two distinct features for the stainless steel which are not found for
pure iron nor the carbon steel. One of this features are the surface structures dis-
played in Fig. 138. The straight lines in a geometrical arrangement with ®xed angles
are slip lines typical for the stress induced movements of dislocations [2].
The other is, that the periodicity found for the meandering treatment of iron in the
surface pro®le produced by the combination of piston e€ect and meandering treat-
ment is not found for the stainless steel, at least it is less pronounced. The shown
e€ect of mechanical material transport due to piston e€ect and meandering is also
the reason, why the saturation with the number of pulses as shown in Chapter 6 is
not e€ective here. Fig. 139 shows the nitrogen integral for laser nitrided stainless

Fig. 138. HRSEM image of the surface of a laser nitrided stainless steel (4 J/cm2, 32 pulses (8  4),
0.7 MPa, homogenized beam).

Fig. 139. Mean nitrogen content in the surface of laser nitrided stainless steel (4 J/cm2, 32 pulses (8  4),
meander).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 141

steel employing the meandering technique. There is no saturation with the number
of pulses, because an increase in the number of pulses here also involves smaller
shifts for subsequent laser spots for the meandering treatment. Finally, it should be
mentioned that no surprises are found in the experiments with isotopic enrichments,
they show the same behavior as for the iron case.
As a summary, the stainless steel behaves during the laser nitriding very much like
iron or carbon steel. The nitrogen take-up and hardness development are almost
identical. Due to the di€erent composition of the material, the phase formation is
di€erent and for the stainless steel there is no formation of martensite or " nitride.
No chromium nitrides were observed after the laser nitriding, in spite of the high
chromium content.

10.3. Comparison with laser nitriding of aluminum

When laser nitriding Al, AlSi and Ti stoichiometric and adherent nitride layers
can form [70]. The formation of AlN by laser nitriding is pressure dependent. Barnikel
gives a threshold pressure of 0.01 MPa above which the AlN fraction grows loga-
rithmically up to the maximum measured pressure of 0.9 MPa [70]. This is explained
by the plasma expansion dynamics. At low pressures the plasma has only short
contact time with the liquid surface, but the time the plasma is acting at the liquid
surface is increasing with the gas pressure. The same piston behaviour imposed by
the laser plasma interaction as for iron was also found in the case of aluminum [70].
The formation of AlN at 4 bar starts at about 0.7 J/cm2 and also increases loga-
rithmically with the laser energy density up to the maximum measured value of 7 J/
cm2. The formation of AlN increases linearly with the number of pulses, but satu-
rates at about 120 pulses. The AlN layer thickness is about 1 mm for 1.5 J/cm2
independent of the number of pulses. It drastically increases for laser ¯uences
exceeding 1.8 J/cm2 and starts to saturate at about 4 J/cm2. The layer thickness
increases with the number of pulses. This increase in the layer thickness was
explained by the piston e€ect also present for Al [70]. For AlSi alloys the formation
of AlN decreases with increasing Si content, but at the same time the surface hard-
ness is increasing.
Fig. 140 shows the nitrogen depth pro®les as obtained via RNRA of laser nitrided
aluminum. At a nitrogen gas pressure of 0.8 MPa and the energy density of 4 J/cm2,
the nitrogen has a maximum of about 30% at the surface, and the pro®le looks not
quite similar to the nitrogen pro®les of laser nitrided iron. For 2 J/cm2 there is also
only a small nitriding e€ect and there is a drop in the nitrogen concentration at the
surface, maybe due to oxidation. At 4 J/cm2 and a nitrogen gas pressure of 10 hPa
there is also a small nitriding e€ect (8 at.%) mainly limited to the surface. Barnikel
and coworkers [70,68,69] are also discussing several mechanism for the nitrogen
take-up:

1. reaction in the plasma and re-condensation afterwards


2. dissolution of nitrogen in the melt along with nitrogen di€usion and phase
formation afterwards
142 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

3. surface reactions with nitrogen forming a thin nitride layer, this layer then
experiences mechanical mixing by melt bath movements and convective ¯ow
with the following pulses.

For the aluminum case they are in favor for the third case but do not exclude a
superposition of all three or even additional mechanisms. The third mechanism
should not be active for the laser nitriding of iron because the isotopic experiments
do not show the thin (50±100 nm) nitride layers which should be clearly visible due
to the high 15 N enrichment. So this mechanism should not be active or it is only a
minor contribution in the iron case.

10.4. Summary of alloys

Additional alloying elements (C, Cr, Ni) do not signi®cantly in¯uence the nitrid-
ing eciency or the nitriding mechanisms during laser treatment of iron based
alloys, at least in those regions where the iron still dominated the properties of the
material. The behavior of `easy nitride formers', as Al or Ti, in connection with laser
nitriding seems to be di€erent. For Al, the addition of Si has tremendous e€ects of
the nitriding eciency and also the pressure dependence of nitriding is di€erent for
Fe and Al.

11. Summary and conclusions

11.1. Summary of raw and homogenized beam irradiations

We ®rst summarize and compare the results obtained for the laser nitriding of iron
with the raw beam and those obtained by the homogenized beam of the excimer

Fig. 140. RNRA nitrogen depth pro®les of laser nitrided pure aluminum (n = 4 pulses, raw beam).
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 143

laser. Only the main di€erences are presented here. More detailed results are given in
the recent paper of Landry et al. [74]. While the homogenized beam can be regarded
as a ¯at-top intensity distribution, the unfocussed raw beam exhibits an intensity
distribution which, in the center of the spot, is higher than the average energy den-
sity. This is the reason for many di€erences between the raw beam and the homo-
genized beam irradiations.
One of these di€erences is the average threshold energy density Hthr . While the
raw beam already invokes a nitriding e€ect starting at H ˆ 1:2 J/cm2, due to the
inhomogeneous intensity distribution, the homogenized beam has a threshold
energy density of 2.4 J/cm2, in agreement with the simulation. This di€erence is also
re¯ected in the larger melting depth dliq for the raw beam treatment.
The piston e€ect, the lateral material transport and the convection are greatly
reduced for the homogenized beam, even almost vanishing, in comparison to the
raw beam irradiations. This fact enabled the separation of some of the e€ects lead-
ing to the saturation in the nitriding. The higher nitrogen concentration hcN i
observed in the spot center for the homogenized beam hinders further nitriding
when it has reached its saturation value, whereas the strong lateral material trans-
port in the case of the raw beam does not allow to reach high saturation values.
Therefore, the take-up of new nitrogen remains constant with an increasing number
of pulses of the raw beam, whereas it saturates for the homogenized beam. In both
cases, the nitrogen depth pro®les can be ®tted with the same superimposed di€usion
pro®les. The saturation concentration is 15 at.% for the homogenized beam and
only 3.4 at.% for its counterpart.
As shown schematically in Fig. 141, no macroscopic material ¯ow is possible if the
temperature and pressure distribution are completely homogeneous across the laser
spot. The inhomogeneous irradiation (a) causes a gradient in temperature, pressure
and melting depth, enabling a pressure-induced expulsion of liquid material at the
outer regions of the laser spot. The homogeneous situation (b) blocks this expulsion
by the overall homogeneous pressure imposed onto the melt. A three-dimensional
simulation including not only the heat and nitrogen di€usion but also the hydro-
dynamic material transport induced by the temperature and pressure gradients
should be able to resolve these phenomena. This simulation was outside the scope of
this work. However it should be pursued in future investigations of the laser nitriding
process.

Fig. 141. Schematic explanation for the presence of a strong piston e€ect only for the inhomogeneous
irradiation.
144 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

As the pressure of the plasma is higher for higher energy densities [270,332], a
large pressure gradient should exist within the plasma above the surface irradiated
by the raw beam. Therefore, a more pronounced lateral expansion of the plasma
plume is expected than in the case of the homogeneous irradiation, leading to a lar-
ger amount of fall-out outside the irradiated area. This lateral expansion leads to
another important di€erence between the laser nitriding with the raw beam and the
homogenized beam, namely, the formation of fall-out deposited outside the laser
spot, which is only observed for the raw beam. This is depicted again in Fig. 142,
which shows a photograph of the spot borders for both cases. The raw beam (a)
exhibits the formation of the ridges as a consequence of the piston e€ect and the
deposition of fall-out outside the spot. The homogeneous irradiation (b) results in a
homogeneous spot with a sharp border. Outside the spot no fall-out can be detected
and only a thermal annealing e€ect around the laser spot is visible, as veri®ed in the
lateral hardness pro®les.
Table 9 again summarizes the results of laser nitriding for both irradiation modes.
Thus, homogenized beam and raw beam laser nitriding both have their pros and
cons. Homogenizing reduces the lateral transport in the molten surface and in the
plasma. This produces rather ¯at nitride coating and enables the separation of the
piston e€ect and convection and allows to study the nitrogen take-up and the
threshold energy density.
The raw beam induces a strong piston e€ect, which can be used to increase the
nitriding depth. The small melting depth of about 1 mm can be increased by the
piston e€ect. By a proper adjustment of the piston e€ect and the laser spot shifts
during the meandering treatment, the nitriding depth can be increased to some 10±
20 mm, thus allowing a large number of technical applications. The meandering of
the laser spot may then also lead to smoother surfaces with an acceptable small
surface roughness.
For both cases, the hardness increase is about the same. It is slightly better for the
raw beam, which also produces a larger hardening depth due to the piston e€ect.
The corrosion resistance should be better with a higher content of the " phase, i.e.
for the raw beam treatment.

11.2. Conclusions and outlook

The investigation of the laser nitriding process is indeed very much like opening
Pandora's box, although here it was only opened partly. Only those changes invoked
by the laser nitriding in the material and detectable after the treatment were inves-
tigated. Diagnosis of the laser±plasma interactions and plasma dynamics during the

Fig. 142. Comparison of the laser spot borders for (a) raw beam and (b) homogenized beam.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 145

nitriding process was not within the scope of this study. Furthermore, the investi-
gations were limited to iron and iron alloys. Nevertheless, the detailed synopsis of
the results has revealed many e€ects occurring during the laser nitriding process and
led to a rather complicated scenario, which was simulated rather successfully by
simple models.
First of all, laser nitriding is related to the laser heating, melting and plasma for-
mation above the liquid surface. The high plasma pressure created by plasma
dynamics enables the dissolution of nitrogen up to high concentrations in the liquid
iron. The threshold in the laser intensity is related to the evaporation of a substantial
amount of material so that a dense plasma and even a plasma absorption wave can
form. The nitrogen is dissolved into the liquid surface in contact with the plasma
and then distributed by di€usion and convection in the material.

Table 9
Comparison of the laser nitriding results for iron obtained with the raw and the homogenized beam

E€ect Quantity Raw beam Homogenized beam


thr 2 h
Threshold in average energy density H 1.2 J/cm 2.4 J/cm2
Phase formation ; 0 ; ; " ; 0 ; ; 00
Melting depth dliq 900±1700 nm 750 nm
Mean nitrogen concentrationa hcN imax 5.0(2) at.% 16.6(5) at.%
n0 1.5(2) 31(2)
Nitrogen pro®les (Eq. (110))b cmax
0 3.4(2) at.% 15.0(6) at.%
n0 6.1(9) 62(9)
cmax
1 17(2) at.% 7(1) at.%
cmax
2 12(1) at.% 20(1) at.%
zmax
1 16(2) nm 33(3) nm
zmax
2 109(3) nm 209(3) nm
Material transporte d‡ 10(1) nm/pulse 0.5(1) nm/pulse
d 23(2) nm/pulse 1.4(2) nm/pulse
Piston e€ect z 55(14) nm/pulse 1.8(4) nm/pulse
Fall-out outside the laser spot Yesf nog
dfo 2(1) nm/pulse 0.0(2) nm/pulse
Material loss dlost 3(2) nm/pulse 0.4(3) nm/pulse
Convection Strong Weak
Surface roughnessc R0a 110(10) nm 15(2) nm
Microhardness Md M0 2.5(3) GPa 2.5(3) GPa
Mmax 4.5(3) GPa 3.5(6) GPa
n0 6.5(8) 56(12)
Hardening depth dh 2mm 1 mm
a
In the topmost 300 nm: hcN i ˆ hcN imax …1 exp… n=n0 ††.
b
Parameters ®tted with activation function when applicable.
p
c
Ra ˆ R0a ‡ R0a n; R0a ˆ 10…5† nm.
d
[HU 0.004] M ˆ M0 ‡ Mmax …1 exp… n=n0 ††.
e
Normalized to the spot size b.
f
Linear with n, depending on spot size, and homogeneity.
g
Fall-out not detectable outside the spot, except for a tiny amount for very large n.
h
This results from the inhomogenities in the spatial intensity distribution. Since this value represents a
mean value, in the central part, the ¯uence of 2.4 J/cm2 is reached already at much lower values of the
average ¯uence for the raw beam.
146 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

There is a saturation in the nitriding e€ect with the number of pulses n and with
the energy density. The saturation with the laser intensity is caused by the increased
evaporation of material and probably also by a transition from the LSC regime to
the LSD or even the LSR regime. Nitriding starts at a nitrogen gas pressure of about
10 hPa, but to achieve high nitriding eciency the pressure should be in the range
from atmospheric pressure up to about 0.7 MPa. In this range, the high nitriding
eciency does not sensitively depend on the pressure. Nevertheless, the results of the
meandering treatments may signi®cantly depend on the pressure because the piston
e€ect and the lateral plasma expansion also depend on it. An optimum of 0.1 MPa
was found for the raw beam/meandering laser nitriding of iron.
The inhomogenity of the spatial laser intensity distribution causes enhanced con-
vection and lateral material transport by the piston e€ect and also an enhanced lat-
eral expansion of the plasma plume above the surface. The latter causes the
deposition of fall-out outside the laser spot. Its amount and the strength of the pis-
ton e€ect and of the convection depend on the spot size and the spatial gradient of
the laser intensity. The fall-out contains " iron-nitride. Neither fall-out nor " phase
are found for a homogeneous irradiation. The e phase is only found after nitriding
iron, when also a strong piston e€ect is active; it probably re-condensates from the
plasma. Nevertheless, the piston e€ect may also be used in combination with the
meandering treatment to achieve thicker nitrided layers, containing the technically
desired " phase.
The pulse duration can also be expected to have a signi®cant in¯uence on the
nitriding e€ect, since it strongly in¯uences the plasma dynamics and the relevant
times available for the nitrogen take-up. Real technical materials, carbon steel and
stainless steel, behave similar to pure iron, except that for stainless steel the " phase
is not formed. The laser nitriding mechanism for iron-based materials may sig-
ni®cantly di€er from that for metals easily forming nitrides (e.g. Ti, Al).
The laser nitriding process can compete with conventional nitriding methods such
as gas nitriding or plasma nitriding. High surface hardness and the formation of the
" phase can be achieved. If the piston e€ect can be used to achieve thick nitrided
layers, many technical applications using the laser nitriding process may arise. Laser
nitriding is a very fast process; it can be accomplished within seconds, and even large
areas can be treated by using lasers with high pulse repetition rates or by using sev-
eral lasers. Nevertheless, laser nitriding is restricted to surfaces with optical accessi-
bility, whereas gas nitriding and plasma nitriding are more ¯exible in this respect.
Furthermore, laser nitriding can also be carried out in air, without the necessity to
use large nitriding chambers, if a thin oxide layer at the surface can be accepted or
protective nitrogen gas ¯ows can be used.
Finally, it can be concluded that laser nitriding is a complex but promising process
composed of several distinct e€ects, which only a combination of complimentary
methods is able to resolve. This work established a ®rst comprehensive investigation
of the process and its dependence on the laser nitriding parameters. A whole set of
parameters (laser ¯uence, number of pulses, spot size, spatial intensity distribution,
pulse duration, wavelength, gas pressure, and material treated) was necessary to
highlight the importance of the various treatment conditions. The in¯uence of these
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 147

parameters was resolved and successfully incorporated into simulations. These


results are a promising basis for further investigations and for process optimization
towards technical applications.
As an outlook, in the following some aspects for further work in this direction are
mentioned which seem to be of special interest and most promising. First, the depth
pro®les should be measured down to greater depth, as well as the convection using
marker layers. This will soon become possible when the new 3 MeV accelerator in
GoÈttingen goes into operation. Plasma diagnostics in connection with a detailed
analysis of the laser-supported absorption waves should bring more information
about the plasma dynamics. The latter can be in¯uenced by the pulse duration which
has also direct e€ects on the piston e€ect and the fall-out deposition. A full theore-
tical simulation of the whole process should be possible, once the in¯uence of all
parameters is known in detail. Therefore, also a variation of the irradiation pattern
should be carried out, maybe in connection with a pilot plant for technical optimi-
zation. Finally, one can also think about additions to the nitrogen atmosphere,
changing the laser±gas interactions.

Acknowledgements

This report is basically my habilitation thesis (Habilitationsschrift) submitted to


the University of GoÈttingen in 1999. I am very much obliged and grateful to Prof.
Dr. Klaus-Peter Lieb for his continuous interest in this work, which he always sup-
ported generously. The long and intensive discussions of the problems that occurred
and his suggestions thoroughly advanced the project and made a large contribution
to the results presented. I also express my deepest gratitude to my co-workers Drs.
Felix Landry, Christof Illgner, Matthias Neubauer, Leena Rissanen, Michael Uhr-
macher, as well as Matthias Niederdrenk and Meng Han. The extent of this report
would be much smaller without their help and their e€orts in carrying out the
experiments and analyzing the results. Thanks are addressed to Detlef Purschke for
expertly operating the IONAS accelerator. I am very grateful to Prof. Dr. Hans
Wilhelm Bergmann, University of Bayreuth, to Drs. Jochen Barnikel, Karsten
Schutte, Andreas Emmel and to Robert Queitsch, all at ATZ-EVUS, Vilseck, for
providing the laser facilities, their help with the laser treatments and for many
enlightening discussions concerning the diculties and problems in laser material
treatments. Most of the laser irradiations presented here were carried out there.
Some laser treatments and pro®lometer measurements were also carried at the Laser
Laboratorium GoÈttingen; thus also the kind cooperation of Prof. Dr. G. Marowsky
and Dr. Klaus Mann is gratefully acknowledged. I also feel obliged to the Institute
of Materials Physics at the University of GoÈttingen, especially to Prof. Dr. R.
Kirchheim, Dr. Hans-Ulrich Krebs and Dieter Plischke for the provision of the
metallographic facilities and the scanning electron microscopy, as well as for many
helpful hints and discussions. I want to thank Prof. Dr Menachem Bamberger,
Technion, Haifa, for his kind hospitality and the provision of his facilities during my
stay there, which was funded by the Max-Planck-Gesellschaft via the MINERVA
148 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

society. The Deutsche Forschungsgemeinschaft (DFG) is gratefully acknowledged


for the ®nancial support (grant Scha 632/3) of this work. Finally, my deepest and
most sincere thanks are going to my wife Ursula and our children Philippe, Ann-
Cathrin, and Leah for their support and their understanding. I have to apologize to
them for the numerous extra hours I spent working instead of being with them.

References

[1] VDI. VDI-Lexikon Werksto€technik. DuÈsseldorf: VDI-Verlag, 1990.


[2] Cahn RW, Haasen P. Physical metallurgy. Amsterdam/Oxford/New York/Tokyo: North-Holland
Physics Publishing, 1983.
[3] Geiger M, Ho€mann P, Deinzer G. Combination of laser beam cutting and welding for the car
body manufacturing: welding with wire as ®ller. In: Bergmann and Kupfer [280], p. 689±99.
[4] Menzies IA, Powell J, Schlenzinger K. Laser beam cutting Ð the state of the art. In: Sossenheimer
and Sepold [279], p. 1±9.
[5] GrundmuÈller F. Laser 1991;3:126±9.
[6] Dorn L. Strahlabsorption beim Laserstrahlpunktschweiûen. In: Sossenheimer and Sepold [279], p.
33±36.
[7] Beyer E. Ein¯uû des laserinduzierten Plasmas beim Schweiben mit CO2-lasern. PhD thesis, TH
Darmstadt, Darmstadt, 1985.
[8] Rosen HG. Using lasers for welding and cutting in industry. In: Mordike [277], p. 319±326.
[9] Amende, W. HaÈrten von Werksto€en und Bauteilen des Maschinenbaus mit dem Hochleis-
tungslaser. DuÈsseldorf Technologie Aktuell 3: VDI-Verlag, 1985.
[10] VDI-Technologiezentrum. Info BoÈrse Laser, 1990, p. 4.
[11] Amende W, Zechmeister H. VDI-Z 1982;124:581±90.
[12] Amende W. Transformation hardening of steel and cast iron with high power lasers. In: Koebner
H, editor. Industrial applications of lasers. Sussex: Wiley, 1984.
[13] Zechmeister H, Amende W. MoÈglichkeiten und Grenzen der Laseranwendungen im Werkzeug-bau.
VDI-Berichte 522, Blechbearbeitung. DuÈsseldorf: VDI-Verlag, 1984. p. 145±62.
[14] Lausch W. Motortechnische Zeitschrift 1985;46:163±9.
[15] Schaaf P, Bauer P, Gonser U. Z Metallkunde 1989;80:77±82.
[16] Bamberger M, Boas M, Akin O. Z Metallkunde 1988;79:806±12.
[17] Amende W. Untersuchung zur Parameterauswahl beim Laserober¯aÈchenveredeln durch Beschich-
ten bzw. Au¯egieren. 1985 Fachinformationszentrum, Karlsruhe BMFT Forschungsbericht T 85-
183.
[18] Schaaf P, Biehl V, Bamberger M, Bauer P, Gonser U. J Mater Sci 1991;26:5019±24.
[19] Steen WM. Surface treatment of materials by laser beams Ð a review. In: Sossenheimer and Sepold
[279], p. 60±4.
[20] Bergmann HW, Breme J, Lee SZ. Laser hardfacing by melt bath reactions. In: Sossenheimer and
Sepold [279], p. 70±3.
[21] Schaaf P, Biehl V, Gonser U, Bamberger M, Langohr M, MaisenhaÈlder F. Hyper®ne Interactions
1990;57:2095±100.
[22] Harmathy P, Nowak G, Amende W. Metallurgical aspects of the production of alloyed skin layers
with the laser beam. In: Mordike [277], p. 83±9.
[23] KuÈpper F, Gasser A, Kreutz EW, Wissenbach K. Cladding of valves with CO2 laser radiation. In:
Bergmann and Kupfer [280], p. 461±7.
[24] Amende W, Nowak G. Hard phase particles in laser processed cobalt rich cladding. In: Bergmann
and Kupfer [280], p. 417±28.
[25] Bergmann HW, Mordike BL. Ober¯aÈchenverglasen durch Laserstrahlschmelzen. Karlsruhe:
Fachinformationszentrum, 1986. BMFT-Forschungsbericht T 86-072
[26] Lin CJ, Spaepen F, Turnbull D. J Non-Cryst Solids 1984;61/62:767±72.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 149

[27] Bergmann HW. Untersuchungen an schnellabgeschreckten, metallischen Systemen. TU Clausthal,


Clausthal-Zellerfeld: Habilitationsschrift, 1983.
[28] BaÈuerle D. Laser processing and chemistry. Springer Series in Materials Science. Berlin, Heidelberg:
Springer Verlag. 1996.
[29] Krebs HU. Int Journal of Non-Equilibrium Processing 1997;10(3):24.
[30] StoÈrmer M, Krebs HU. J Appl Phys 1996;78:7080±7.
[31] Krebs HU, Luo Y, StoÈrmer M, Crespo A, Schaaf P, Bolse W. Appl Phys A 1995;61:591±4.
[32] StoÈrmer M, Crespo A, Schaaf P, FaÈhler S, Krebs HU. Nucl Instrum and Methods 1997;B122:503±
6.
[33] Kautek W, Pentzien S, Conradi A, Kruger J, Brzezinska KW. Appl Surf Sci 1996;106:158±65.
[34] Proyer S, Stangl E, Borz M, Hellebrand B, BaÈuerle D. Physica C 1996;257:1±15.
[35] Verbist K, KuÈhle A, Vasiliev AL. Physica C 1996;269:131±8.
[36] D'Anna E, De Giorgi ML, Leggieri G, Luches A, Martino M, Perrone A, Mihailescu IN, Mengucci
P, Drigo AV. Thin Solid Films 1992;213:197±204.
[37] Chen XY, Xiong SB, Sha ZS, Liu ZG. Appl Surf Sci 1997;115:279±84.
[38] Schmidt H, Ihlemann J, Wol€-Rottke B, Luther K, Troe J. J Appl Phys 1998;83:5458±68.
[39] Preuss S, Demchuk A, Stuke M. Appl Phys 1995;A61:33±7.
[40] Zergioti I, Stuke M. Appl Phys 1998;A67:391±5.
[41] Schaaf P, Emmel A, Schubert E, Bergmann HW, Lieb KP. Hyper®ne Interactions 1994;92:1361±6.
[42] Schaaf P, Emmel A, Illgner C, Lieb KP, Schubert E, Bergmann HW. Mater Sci Eng A
1995;197:L1±0L4.
[43] Emmel A. Aufbau und Eigenschaften von Ober¯aÈchen®lmen nach einer Excimerlaserbelichtung
und deren Bedeutung fuÈr das Prozeûverhalten. Ph.D. thesis, UniversitaÈt Erlangen-NuÈrnberg, 1993.
[44] Schaaf P, Illgner C, Niederdrenk M, Lieb KP. Hyper®ne Interactions 1995;95:199±225.
[45] Schaaf P. Hyper®ne Interactions 1998;111:113±9.
[46] Schaaf P, Illgner C, Landry F, Lieb KP. Surface and Coatings Technology 1998;100(101):399±402.
[47] Schaaf P, Landry F, Neubauer M, Lieb KP. Hyper®ne Interactions 1998;113:429±34.
[48] Schaaf P, Landry F, Lieb KP. Appl Phys Lett 1999;74(1):153±5.
[49] Schaaf P, Landry F, Lieb KP. Laser nitriding and ion beam analysis. In: Duggan JL, Morgan IL,
editors. Proc. International Conference on the Applications of Accelerators in Research and
Industry, ICAARI 1998, Denton, Texas, volume AIP CP 465, AIP Conference Proceedings, 1999.
p. 737±40.
[50] Illgner C. Untersuchungen zum Lasernitrieren von Eisen. Ph.D. thesis, UniversitaÈt GoÈttingen,
GoÈttingen, 1997.
[51] Illgner C, Schaaf P, Lieb KP, Schubert E, Queitsch R, Bergmann HW. Appl Phys A 1995;61:1±5.
[52] Illgner C, Lieb KP, Schaaf P, Mann K, KoÈster H, Marowsky G. Appl Phys A 1996;62:231±6.
[53] Illgner C, Schaaf P, Lieb KP, Queitsch R, Schutte K, Bergmann HW. Nucl Instrum and Methods
1997;B122:420±2.
[54] Illgner C, Schaaf P, Lieb KP, Queitsch R, Schutte K, Bergmann HW. Appl Surf Sci 1996;109/
110:150±3.
[55] Illgner C, Schaaf P, Lieb KP, Queitsch R, Barnikel J. J Appl Phys 1998;83:2907±14.
[56] Landry F, Neubauer M, Lieb KP, Schaaf P. Optimised irradiation parameters for laser nitriding of
iron and X5CrNiMo18.10. In: Mordike BL, editor. Proc. European Conference of Laser Treatment
of Materials, EKLAT 1998, September 1998, Hannover. Frankfurt: Werksto€-Informationsge-
sellschaft, 1998. p. 81±6.
[57] Landry F, Schaaf P, Neubauer M, Lieb KP. Appl Surf Sci 1999;138(139):266±70.
[58] Ursu I, Mihailescu IN, Nanu L, Nistor LC, Popescu M, Teodorescu VS, Prokhorov AM, Konov
VI, Uglov SA, Ralchenko VG. J Phys D 1986;19:1183±8.
[59] Galiev AL, Krapivin LL, Mirkin LI, Uglov AA. Sov Phys Dokl 1980;25:208.
[60] Uglov AA, Ignatev MB, Gnedovets AG, Smurov IY. J de Physique 1989;50:C5±727.
[61] Uglov AA, Nizametdinov MM. Fizika i Khimiya Obrabotki Materialov 1977;2:133.
[62] Uglov AA, Galiev AL. Fizika i Khimiya Obrabotki Materialov 1978;3:23.
[63] Drobnik A, Rozniakowski K, Jablonski W. Materials Research Bulletin 1979;18:1049.
150 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

[64] Fastow M, Bamberger M. Scripta Met 1988;22:183±6.


[65] Bergmann HW. Z Werksto€technik 1985;16:392.
[66] Nakamura K, Hikita M, Asano H, Terada A. Jpn J Appl Phys 1982;21:672.
[67] Ogalc SB, Patil PP, Phase DM, Bhandarkar YV, Kulkarni SK, Kulkarni S, Ghaisas SV, Kanetkar
SM, Bhide VG, Guha S. Phys Rev B 1987;36:8237±50.
[68] Barnikel J, Schutte K, Bergmann HW. HaÈrterei-Technische Mitteilungen 1997;52:91±3.
[69] Barnikel J, Seefeld T, Schutte K, Bergmann HW. HaÈrterei-Technische Mitteilungen 1997;52:94±6.
[70] Barnikel J. Nitrieren von Aluminiumwerksto€en mit UV-Laserstrahlung. Ph.D. thesis, UniversitaÈt
Erlangen-NuÈrnberg, 1998.
[71] Meneau C, Andreazza P, Andreazza-Vignolle C, Gideau P, Villain JP, Boulmer-Leborgne C. Sur-
face and Coatings Technology 1998;100/101(1±3):12±16.
[72] Laude LD, Kolev K, Brunel M, Deleter P. Appl Surf Sci 1995;86:368±81.
[73] Troe J. An investigation of the laser nitriding process is quite similar to the opening of Pandora's
box. Private communication, 1993.
[74] Landry F, Lieb KP, Schaaf P. J Appl Phys 1999;86:168±78.
[75] Landry F. Lasernitrieren von Armco-Eisen und Eisenwerksto€en. Ph.D. thesis, UniversitaÈt GoÈt-
tingen, GoÈttingen, 1999.
[76] Lieb KP, Bolse W, Corts T, MuÈller W, Osipowicz T, Weber T. Nucl Instr and Meth 1990;B50:10.
[77] Corts T, Bolse W, Osipowicz T, Lieb KP. Appl Phys 1990;A51:537.
[78] Kacsich T, Niederdrenk M, Schaaf P, Lieb KP, Geyer U, Schulte O. Surface and Coatings Tech-
nology 1997;93:32±6.
[79] ASM. Metals handbook. American Society for Metals, Metals Park, OH, 1978.
[80] Daeves K In: Verein Deutscher EisenhuÈttenleute, editor. Werksto€-Handbuch Stahl und Eisen.
DuÈsseldorf: Verlag Stahleisen, 1965.
[81] Brooks CR. Principles of the surface treatment of steels. Lancaster, Basel: Technomic Publishing
Company, 1992.
[82] Lai WY, Zheng QC, Hu WY. J Phys: Condens Matter 1994;6L:259.
[83] Sawada H, Nogami A, Matsumiya T. Phys Rev 1994;B50:10004.
[84] Jack KH. J Alloys and Compounds 1995;222:160.
[85] Husang MZ, Ching WY. Phys Rev 1995;B51:3222.
[86] Weber T, de Wit L, Saris FW, Schaaf P. Thin Solid Films 1996;279:217±21.
[87] Takahashi H, Komuro M, Hiratani M, Igarashi M, Sugita Y. J Appl Phys 1998;84(3):1493±8.
[88] Keramikverbund Karlsruhe-Stuttart. Hochleistungskeramik: von Pulvern zu Bauteilen. DuÈsseldorf:
VDI-Verlag, 1992.
[89] Terai T. Surface and Coatings Technology 1998;106(1):18±22.
[90] Savart M. Ann Chim et Phys (2) 1828;37:326.
[91] HaÈgg G. Nature 1928;121:826.
[92] HaÈgg G. Nature 1928;122:314.
[93] Fry A. Stahl Eisen 1923;43:1271.
[94] Fry A. Kruppsche Monatshefte 1923;4:137.
[95] Lehrer E. Z Electrochem 1930;36:383.
[96] Lehrer E. Z Electrochem 1930;36:460.
[97] Jack KH. Proc Roy Soc 1948;A195:34.
[98] Jack KH. Proc Roy Soc 1948;A195:41.
[99] Jack KH. Acta Cryst 1950;3:392.
[100] Jack KH. Proc Roy Soc 1951;A208:200.
[101] Jack KH. Proc Roy Soc 1951;A208:216.
[102] Jack KH. Acta Crystallogr 1952;5:404.
[103] Jack DH, Jack KH. Mater Sci Eng 1973;11:1±27.
[104] Eisenhut O, Kaupp E. Z Elektrochem 1930;36(6):392±404.
[105] Paranjpe VG, Cohen M, Bever MB, Floc CF. Trans AIME 1950;188:261±7.
[106] Wriedt HA, Gokcen NA, Nafziger RH. Bulletin of Alloy Phase Diagrams 1987;8:355±77.
[107] Kunze J. Nitrogen and carbon in iron and steel. Berlin: Akademie Verlag, 1990.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 151

[108] Guillermet AF, Du H. Z Metallkunde 1994;85:154.


[109] Kaufman L, Bernstein H. Computer calculation of phase diagrams. New York: Academic Press,
1970. [see also the CALPHAD journal, Pergamon Press]
[110] Frisk K. Calphad 1991;15:79±106.
[111] Hirschfeld D., editor. Thermodynamik und Ordnung von Sticksto€atomen in Eisennitriden:
Konsequenzen fuÈr das Fe±N Zustandsschaubild. DGM Informationsgesellschaft, 1995.
[112] Kooi BJ, Somers MAJ, Mittemeijer EJ. Metall Mater Trans 1996;A27:1063±71.
[113] Kooi BJ, Somers MAJ, Mittemeijer EJ. Thermodynamik der Fe±N Phasen und das Fe±N
Zustandsschaubild. In: Mittemeijer and Grosch [198], p. 9±18.
[114] Kunze J. HaÈrterei-Technische Mitteilungen 1996;51:348±55.
[115] Kunze J, Broz P, Stloukal I. Steel Research 1996;67(7):279±84.
[116] Hertzman S, Jarl M. Metall Trans 1987;A18(10):1745±52.
[117] Frisk K. Metall Trans 1990;A21(9):2477±88.
[118] Gabidullin RM, Kolachev BA, Krasnova EV. Izv. VUZ Chernaya Metall. 1980;7:5±9. Translation:
Steel USSR 1980;7:348±350.
[119] Wriedt HA, Gokcen NA,, Nafziger RHMassalski TBBinary alloy phase diagrams 1990;vol. 2:1728±
30American Society for Metals (ASM)Metals Park, OH.
[120] Kunze J. Steel Research 1987;57:361±7.
[121] Hillert M, Jarl M. Metall Trans 1975;A6:553±9.
[122] AÊgren J. Metall Trans 1979;A10:1847±52.
[123] Du H. J Phase Equilibria 1993;14(6):682±93.
[124] Hillert M, Sta€anson LI. Acta Chem Scand 1970;24:3618±26.
[125] Harvig H. Acta Chem Scand 1971;25:3199±204.
[126] Kikuchi R, de Fontaine D. National Bureau of Standards Special Publication 1977;496:967±98.
[127] Mohri T, Sanchez JM, de Fontaine D. Acta Metall 1985;33:1171±85.
[128] Kooi BJ. Iron-nitrogen phases: thermodynamics, long range order and oxidation behaviour. Ph.D.
thesis, Technische Universiteit Delft, Delft, 1995.
[129] Kooi BJ, Somers MAJ, Mittemeijer EJ. Metall Mater Trans 1994;25A:2797±814.
[130] Kooi BJ, Somers MAJ, Mittemeijer EJ. Metall Mater Trans 1996;A27:1055±61.
[131] Horstmann D. Das Zustandsschaubild Eisen-Kohlensto€. DuÈsseldorf: Verlag Stahleisen mbH,
1985.
[132] Schumann H. Metallographie. VEB Deutscher Verlag fuÈr Grundstondustrie, Leipzig, 1987.
[133] Jacobs H, Rechenbach D, Zachwieja U. J Alloys and Compounds 1995;227(1):10±17.
[134] Rechenbach D, Jacobs H. J Alloys and Compounds 1996;235(1):15±22.
[135] Srinivasan A, Reynders B, Grabke HJ. Steel Research 1995;66:439±43.
[136] Weber T, de Wit L, Saris FW, KoÈniger A, Rauschenbach B, Wolf GK, Krauss S. Mater Sci Eng
1995;A199:205.
[137] Rochegude P, Foct J. Phys Stat Sol (A) 1985;88:137±42.
[138] Rochegude P, Billon B. Scripta Metall 1986;20:1095.
[139] Hendricks SB, Kosting PR. Z Kristallogr 1930;74:511.
[140] Rechenbach D. Das System Eisen/Sticksto€: Strukturen und physikalische Eigenschaften von
Eisennitriden. Ph.D. thesis, UniversitaÈt Dortmund, Dortmund, 1995.
[141] Somers MAJ, Kooi BJ, Maldzinski L, Mittemeijer EJ, van der Horst AA, van der Kraan AM, van
der Pers NM. Acta Metall Mater 1997;45:2013±25.
[142] Pekelharing MI, BoÈttger A, Somers MAJ, Steenbergen MP, van der Kraan AM, Mittemeier EJ.
Proc. High Nitrogen Steels 1998. HaÈnninen H, Hertzman S, Romn J, editors. Materials Science
Forum 1999;318±320:115±20.
[143] Heiman N, Kazama NS. J Appl Phys 1981;20:492.
[144] Oueldennaoua A, Bauer-Grosse E, Foos M, Frantz C. Scripta Metall 1985;19:1503.
[145] Jiang E, Sun C, Li J, Liu Y. J Appl Phys 1989;65:1659.
[146] Suzuki K, Morita H, Kaneko T, Yoshida H, Fujimori H. J Alloys and Compounds 1993;201:232.
[147] Rissanen L, Lieb KP, Niederdrenk M, Schulte O, Schaaf P. Structure and stability of the new cubic
FeN phase. In: Ortalli I, editor. Proc. Internat. Conf. Applications MoÈssbauer E€ect (ICAME),
152 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

Sept. 10±16, 1995, Rimini, Italy, Conference Proceedings SIF, vol. 50. Bologna: Compositori, 1996.
p. 595±8.
[148] Rissanen L, Neubauer M, Lieb KP, Schaaf P. J Alloys and Compounds 1998;274(1/2):74±82.
[149] Rissanen L, Schaaf P, Neubauer M, Lieb KP, Keinonen J, Sajavaara T. Appl Surf Sci 1999;138/
139:261±5.
[150] Hinomura T, Nasu S. 57Fe MoÈssbauer study of Fe nitrides. In: Ortalli I, editor. Proc. Internat.
Conf. Applications MoÈssbauer E€ect (ICAME), Sept. 10±16. Rimini, Italy: Compositori SIF, 1995.
[151] Kume W, Yamagishi H. J Phys Soc Jpn 1964;19:414.
[152] Nakagawa H, Nasu S, Fujii H, Takahashi M, Kanamaru F. Hyper®ne Interactions 1991;69:455±8.
[153] Hinomura T. MoÈssbauer study of Fe±N alloys and compounds (in English). Ph.D. thesis, Osaka
University, Division of Materials Physics, Osaka, 1998.
[154] Schenck H, Frohberg MG. Arch EisenhuÈttenwesen 1962;33:593±600.
[155] Feichtinger H, Satir-Kolorz A, Xiao-hong Z. Solubility of nitrogen in solid and liquid iron alloys
with special regard to the melting range. In: Proceedings High Nitrogen Steels 1988. London: The
Institute of Metals, 1989.
[156] Lakomski VI, Torkhov GF. Sov Phys Dokl 1969;13:1159.
[157] Somers MAJ, Mittemeijer EJ. Metall Trans 1989;A20:1533±9.
[158] Micrometrics GmbH. Keramik-Ing. 1993;4:55.
[159] Gonser U, Meechan CJ, Muir AH, Wiedersich H. J Appl Phys 1963;34:2373±8.
[160] Mekata M, Yoshimura H, Takaki H. J Phys Soc Japan 1972;33:62±9.
[161] Yamaoka T, Mekata M, Takaki H. J Phys Soc Jpn 1973;35:63±7.
[162] Bainbridge J, Channing DA, Whitlow WH, Pendlebury RE. J Phys Chem Solids 1973;34:1579±86.
[163] DeChristofaro N, Kaplow R. Metall Trans A 1977;8:425±30.
[164] Fromm E, Jehn H, Hehn W, Speck H, HoÈrz G. Gases and carbon in metals (thermodynamics,
kinetics, and properties): iron±nitrogen, vol. XV of Physik Daten. FIZ Energie, Physik, Mathema-
tik, Karlsruhe, 1982.
[165] Somers MAJ, Mittemeijer EJ. Metall Mater Trans 1995;A26:57±74.
[166] Torchane L, Bilker P, Dulcy J, Gantois M. Report, Laboratoire de Science et Genie des Surfaces,
Nancy, 1994.
[167] Somers MAJ, Mittemeijer EJ. Modelling of the kinetics of nitriding and nitrocarburizing of iron.
In: Milam DL, Poteet DA, Pfa€mann GD, Rudnev V, Muehlbauer A, Albert WB, editors. Proc.
17th ASM Heat Treating Society Conference, 1998. p. 321±30.
[168] Schaaf P, Bauer P, Gonser U. Hyper®ne Interactions 1989;46:541±8.
[169] Schaaf P, Biehl V, Bamberger M, Sha®rstien G, Langohr M, MaisenhaÈlder F, Gonser U. Laser
surface alloying of steel with CrB2. In: Bamberger M, Schorr M, editors. Proceedings 5th Israel
Materials Engineering Conference, December 19±20, 1990, Haifa, Israel. London/Tel Aviv: Freund
Publishing House, 1991. p. 451±4.
[170] Ariely S, Bamberger M, HuÈgel H, Schaaf P. Solidi®cation and phase transformation in AISI 1045
steel laser surface alloyed with TiC. In: Denney P, Miyamato I, Mordike BL, editors. Proceedings
ICALEO'93, October 1993, Orlando, FL, vol. 77, 1994.
[171] Ariely S, Bamberger M, HuÈgel H, Schaaf P. J Mater Sci 1995;30:1849±53.
[172] Liedtke D. Nitrieren. Mcrkblatt 447. VdEh, DuÈsseldorf, 1974.
[173] Kunst H, Liedtke D. Badnitrieren von Eisenwerksto€en Ð Untersuchungen an Nitridschichten,
vol. 7 of Tribologic, Reibung Ð Verschleiss Ð Schmierung, 1983, p. 39.
[174] Boyer H, editor. Case hardening of steel. Metals Park, OH: ASM International, 1987.
[175] Abada L, Rixecker G, Aubertin F, Schaaf P, Gonser U. Phys Stat Sol (A) 1993;139:181±7.
[176] Williamson DL, Ozturk O, Glick S, Wei R, Wilbur PJ. Nucl Instrum Methods B 1991;59/60:737±
41.
[177] Marest G. Defect and Di€usion Forum 1988;57-58:273±326.
[178] Niederdrenk M, Schaaf P, Lieb KP, Schulte O. J Alloys and Compounds 1996;237:81±8.
[179] Bell T. HaÈrterei-Technische Mitteilungen 1985;30:161±4.
[180] Cherry FK. Heat Treatment of Metals 1987;14:1±5.
[181] MaÈdler K, Bergmann W, Dengel D. HaÈrterei-Technische Mitteilungen 1996;51:338±47.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 153

[182] Wahl G. Fachber HuÈttenpraxis Metallweiterverarbeitung 1981;19:1076.


[183] Dawes C, Tranter DF. Heat Treat Met 1985;12:70.
[184] Juza R. Adv Inorg Chem Radiochem 1967;9:81.
[185] Burdese A. Metall Italiana 1957;3:195.
[186] Somers MAJ, Mittemeijer EJ. HaÈrterei-Technische Mitteilungen 1987;42:321±31.
[187] Kooi BJ, Somers MAJ, Jutte RH, Mittemeijer EJ. Oxidation of Metals 1997;48:111±28.
[188] Wagner G. Untersuchung der ober¯aÈchennahen Schichten von stickstomplantiertem Stahl. Ph.D.
thesis, UniversitaÈt des Saarlandes, SaarbruÈcken, 1990.
[189] Vredenberg AM, Perez-Martin CM, Custer JS, Boerma DO, de Wit L, Saris FW, van der Pers NM,
de Keijser TH, Mittemeijer EJ. J Mater Res 1992;7(10):2689±712.
[190] Pant P, Dahlmann P, Schlump W, Stein G. Steel Research 1987;58(1):18.
[191] Kumoro M, Kozono Y, Hanazono M, Sugita Y. J Appl Phys 1990;67:5126±30.
[192] Crank J. The mathematics of di€usion. Oxford: Clarendon Press, 1975.
[193] Kirchheim R, Fromm E. High Temperatures Ð High Pressures 1979;6:329±39.
[194] Kirchheim R. Physica Status Solidi (B) 1979;91:123±34.
[195] Mathuni J, Kirchheim R, Fromm E. Scripta Metall 1979;13:631±4.
[196] Verein Deutscher EisenhuÈttenleute, editor. Werksto€kunde Stahl, Band 1: Grundlagen. Berlin/
Heidelberg/New York: Springer Verlag, 1984.
[197] Mittemeijer EJ, Grosch J, editors. Berichtsband Nitrieren und Nitrocarburieren 1991. Wiesbaden:
AWT, 1991.
[198] Mittemeijer EJ, Grosch J, editors. Berichtsband Nitrieren und Nitrocarburieren 1996. Wiesbaden:
AWT, 1996.
[199] Barin I, Knacke O. Thermochemical properties of inorganic substances. Berlin±Heidelberg±New
York: Springer Verlag, 1973. [also Verlag Stahleisen, DuÈsseldorf]
[200] Ho€mann F, Ho€mann R, Mittemeijer EJ. HaÈrterei-Technische Mitteilungen 1992;47:365.
[201] Somers MAJ, Mittemeijer EJ. Kinetik des Nitrierens und Nitrocarburierens: Verbindungs-
schichtwachstum. In: Mittemeijer and Grosch [198], 1996. p. 157±66.
[202] Grabke HJ. Ber Bunsenges Phys Chem 1968;72:533±40.
[203] Somers MAJ, Mittemeijer EJ. Verbindungsschichtbildung waÈhrend Gasnitrieren und Gas- und
Salzbadnitrocarburieren. In: Mittemeijer and Grosch [197], p. 24±38.
[204] Somers MAJ, Mittemeijer EJ. Materials Science Forum 1992;102-104:223±8.
[205] Dimitrov VL, D'Haen J, Knuyt G, Quaeyhaegens C, Stals LM. Appl Phys 1996;A63:475±80.
[206] Dimitrov VI, Knuyt G, Stals LM, D'Haen J, Quaeyhaegens C. Appl Phys 1998;A67:183±92.
[207] Kofstad P. High temperature oxidation of metals. New York: Wiley, 1966.
[208] Birks N, Meier GH, Pettit FS. In: High temperature corrosion. Superalloys, supercomposites and
superceramics. New York: Academic Press, 1989. p. 439±89.
[209] Pietzsch S, BoÈhmer S. HaÈrterei-Technische Mitteilungen 1996;51:364±71.
[210] Somers MAJ, Mittemeijer EJ. Eigenspannungen in der Verbindungsschicht. In: Mittemeijer and
Grosch [197], p. 125±138.
[211] Spies HJ, Schaaf P, Vogt F. Materialwissenschaft und Werksto€technik 1998;29:588±94.
[212] Mittemeijer EJ, Somers MAJ. Surface Engineering 1997;13:483±97.
[213] Ho€mann R, Mittemeijer EJ, Somers MAJ. HaÈrterei-Technische Mitteilungen 1994;49:177.
[214] Berg HJ, Spies HJ, BoÈhmer S. HaÈrterei-Technische Mitteilungen 1991;46:375.
[215] Spies HJ, HoÈck K. HaÈrterei-Technische Mitteilungen 1996;51:233±7.
[216] Somers MAJ, Mittemeijer EJ. Oxidschichtbildung und gleichzeitige GefuÈgeaÈnderung der Verbin-
dungsschicht. In: Mittemeijer and Grosch [197], p. 152±62.
[217] Spies HJ, Vogt F. HaÈrterei-Technische Mitteilungen 1997;52:342±9.
[218] Spies HJ, Bergner D. HaÈrterei-Technische Mitteilungen 1992;47:346±56.
[219] Qiang YH. Wear 1998;218(2):232±6.
[220] O'Brian JM, Goodman D. Heat Treatment 1991;4:420.
[221] Rie KT. Schram DC. Stals L. Wolf GK. editors. Surface and coatings technology. vol. 74/75, 1995.
[222] Matthews A. Sartwell BD. editors. Surface and coatings technology. vol. 85, 1996.
[223] GuÈnzel R, MoÈller W, Wieser F, editors. Surface and coatings technology. vol. 93, 1997.
154 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

[224] Richard A, Kassing R, Pulker H, Rie KT, editors. Surface and coatings technology. vol. 97/98,
1998.
[225] Riviere JP, Pranevicius L, Martinez-Duart JM, Grill A, editors. Surface and Coatings Technology,
1998. vol. 100/101, E-MRS Conference 1997.
[226] Perriere J, Boyd IW, Stuke M, Biermann U, editors. Appl Surf Sci, E-MRS Conference 1998.
[227] Edenhofer B. Heat Treat Mat 1974;1:27.
[228] Marciniak A. Surf Eng 1985;1:283.
[229] Metin E, Inal OT. J Mater Sci 1987;22:2783.
[230] Meletis EI, Yan S. J Vac Sci Technol 1993;A 11:25.
[231] Dimitrov VI, D'Haen J, Knuyt G, Quaeyheagens C, Stals LM, A model describing surface mod-
i®cation of metals by plasma. 1999 (in press).
[232] Wagner C. Z Phys Chem 1933;21B:25.
[233] Seith W. Di€usion in metallen. Berlin, Heidelberg: Springer Verlag, 1955.
[234] Kacsich T, Lieb KP. Thin Solid Films, Letters 1994;245:4±6.
[235] Kacsich T, Neubauer M, Geyer U, Baumann K, Rose F, Uhrmacher M. J Phys: Cond Matter
1995;28:424.
[236] Kelly R. Appl Surf Science 1998;133(4):251±69.
[237] Chowdhury K, Vispute RD, Jagannadham K, Narayan J. J Mater Res 1996;11(6):1458±69.
[238] Wagner G, Leutenecker R, Gonser U. Hyper®ne Interactions 1990;56:1653±6.
[239] Rauschenbach B, Kolitsch A. Phys stat sol (a) 1983;80:211±22.
[240] Rauschenbach B, Kolitsch A. Phys stat sol (a) 1983;80:471±82.
[241] Antilla A, Keinonen J, Uhrmacher M, VahvaselkaÈ S. J Appl Phys 1985;57:1423.
[242] Piette M, Terwagne G, MoÈller W, Bodart F. Mater Sci Eng 1989;B 2:189±94.
[243] Terwagne G, Piette M, Bertrand P, Bodart F. Mater Sci Eng 1989;B 2:195±201.
[244] Leutenecker R, Wagner G, Louis T, Gonser U, Guzman L. Mater Sci Eng 1989;A 115:229.
[245] Chabica ME, Williamson DL, Wei R, Wilbur PJ. Surf Coat Technol 1992;51:24.
[246] Williamson DL, Wang L, Wei R, Wilbur PJ. Mater Lett 1990;9:302.
[247] Ozturk O, Williamson DL. J Appl Phys 1995;77:3839.
[248] Wei R. Surface and Coatings Technology 1996;83:218.
[249] Wei R, Vajo JJ, Matossian JN, Wilbur PJ, Davies JA, Williamson DL, Collins GA. Surface and
Coatings Technology 1996;83:235.
[250] Wei R, Shogrin B, Wilbur PJ, Ozturk O, Williamson DL. J Tribol 1994;116:870.
[251] Williamson DL, Davies JA, Wilbur PJ. Surface and Coatings Technology 1998;103-104(1-4):178±84.
[252] Williamson DL, Davies JA, Wilbur PJ, Vajo JJ, Wei R, Matossian JN. Nucl Instrum and Methods
1997;B 127/128:930±4.
[253] Ziegler JF. Biersack JP. Littmark U. PC program package TRIM95. 1995.
[254] Scapellato N, Uhrmacher M, Lieb KP. J Phys F: Metal Phys 1988;18:677.
[255] Kehrel A, Lieb KP, Uhrmacher M. Mater Sci Eng 1989;A115:43.
[256] Weber T, Lieb KP. Nucl Instr and Meth 1989;B 44:54.
[257] Mantese JV, Brown IG, Cheung NW, Collins GA. Mater Res Soc Bull 1996;21(8):52.
[258] Samandi M, Shedden BA, Smith DI, Collins GA, Hutchings R, Tendys J. Surface and Coatings
Technology 1993;59:261.
[259] Lei MK, Zhang ZL. J Vac Sci Technol 1997;A 15:421.
[260] Williamson DL, Ozturk O, Wi R, Wilbur PJ. Surface and Coatings Technology 1994;65:15.
[261] Williamson DL. Nucl Instrum and Methods 1993;B76:262±7.
[262] Conrad JR. J Appl Phys 1987;62:777.
[263] Ensinger W. Surface and Coatings Technology 1998;100-101(1-3):336±47.
[264] Ensinger W, HoÈchbauer T, Rauschenbach B. Surface and Coatings Technology 1998;103-104(1-4):
218±21.
[265] Blawert C, Mordike BL, Collins GA, Short K. Surface and Coatings Technology 1998;103-104(1-4):
240±7.
[266] Ready JF. J Appl Phys 1965;36:462±8.
[267] Ready JF. E€ects of high-power Laser radiation. New York±London: Academic Press, 1971.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 155

[268] von Allmen M. J Appl Phys 1976;47:5460.


[269] von Allmen M In: Laser beam interactions with materials. Springer Series in Materials Science,
vol. 2. Berlin, Heidelberg, New York: Springer Verlag, 1987.
[270] Raizer YP. Sov Phys JETP 1970;31:1148.
[271] BaÈuerle D. Chemical processing with Lasers. Springer Series in Materials Science Berlin, Heidel-
berg: Springer Verlag, 1986.
[272] von Allmen M, Blatter A. Laser-beam interactions with materials. Springer Series in Materials
Sciences Berlin: Springer Verlag, 1983.
[273] Einstein A. Physikal Zeitschrift 1917;18:121±8.
[274] Maimann TH. Nature 1960;187:493±4.
[275] Bertolotti M, editor. Physical processes in laser-materials interactions. New York±London: Plenum
Press, 1983.
[276] Laser materials processing. In: Bass M., editor. volume 3 of Materials processing: theory and
practices. Amsterdam, New York, Oxford: North-Holland, 1983.
[277] Mordike BL. Laser treatment of materials, Proc. 1st European Conference on Laser Treatment of
Materials, ECLAT 1986, Bad Nauheim, Germany: DGM Informationsgesellschaft, Oberursel. 1987
[278] Waidelich W, editor. Laser, Optoelektronik in der Technik, Proc. Eighth International Congress
`LASER87', MuÈnchen, Germany. Berlin, Heidelberg, New York: Springer Verlag, 1987.
[279] Sossenheimer H, Sepold G, editors. Proc. 2nd European Conference on Laser Treatment of Mate-
rials, ECLAT 1988. Bad Nauheim, Germany: DVS-Verlag, DuÈsseldorf; Berlin, Heidelberg, New
York: Springer Verlag, 1988.
[280] Bergmann HW, Kupfer R. Proc. 3rd European Conference on Laser Treatment of Materials,
ECLAT 1990, Erlangen, Germany: Sprechsaal Publishing Group, Coburg, 1990.
[281] Gobrecht H. Optik, vol. 3, Bergmann-SchaÈfer: Lehrbuch der Experimentalphysik. Berlin, New
York: Verlag Walter de Gruyter, 1978.
[282] Tradowski K. Laser: Grundlagen, Technik, Anwendungen. WuÈrzburg: Vogel-Verlag, 1979.
[283] Weber H, Herziger G. Laser: Grundlagen und Anwendungen. Weinheim: Physik-Verlag, 1979.
[284] KneubuÈhl FK, Sigrist MW. Laser. Stuttgart: Teubner TaschenbuÈcher Physik, 1995.
[285] Bass, M., Lasers for Laser materials processing. In: Bass [276], chapter 1, p. 1±14.
[286] Spalding IJ. Characteristics of Laser beams for machining. In: Bertolotti [275], p. 1±47.
[287] Manzur T, de Maria T, Chen W, Roychoudhuri C. Proc SPIE 1996;2703:490±501.
[288] Bergmann HW, Lee SZ. Opto Elektronik Magazin 1987;3:623±30.
[289] Sowada U, Kahler HJ, Basting D. World Lasers Alamanac 1988;1:50±1.
[290] Schmatjko KJ, Endres G. World Lasers Alamanac 1988;1:46±9.
[291] CRC. Handbook of chemistry and physics, 63rd ed., Boca Raton: CRC Press, 1982/83.
[292] Bettis JR. Appl Opt 1992;31:3448±52.
[293] Weyl GM. Physics of Laser-induced brcakdown: An update. In: Radziemski LJ, Cremers DA,
editors. Laser-induced plasmas and applications. New York: Marcel Dekker, 1992.
[294] Allmen MF. Coupling of Laser radiation to metals and semiconductors. In: Bertolotti [275], chapter
2, p. 49±75.
[295] Marowsky G, Rhodes CK. Appl Phys 1998;B 66:475±8.
[296] Grosse P. Freie Elektronen in FestkoÈrpern. Berlin: Springer Verlag, 1979.
[297] Schulze GER. Metallographie. Wien, New York: Springer Verlag, 1983.
[298] Jackson JD. Klassische Elektrodynamik. Berlin, New York: Verlag Walter de Gruyter, 1983.
[299] Gregson VG. Laser heat treatment. In: Bass [276], chapter 4, p. 201±233.
[300] Hellwege KH. Madelung O. Landolt-BoÈrnstein, vol. III 15 b of Neue Serie. 1985.
[301] Touloukian YS, Buyco EH Speci®c heat, metallic elements and alloys. In: Touloukian YS, Ho CY,
editors. Thermophysical properties of matter, vol. 4. New York: IFI/Plenum, 1983.
[302] Touloukian YS, Powell RW, Ho CY, Klemens PG. In: Touloukian YS, Ho CY, editors. Thermal
conductivity, metallic elements and alloys, vol. 1. Thermophysical properties of matter. New York:
IFI/Plenum, 1970
[303] Touloukian YS, Powell RW, Ho CY, Nicolaou MC. Thermal di€usivity. In: Touloukian YS, Ho
CY, editors. Thermophysical properties of matter, vol. 10. New York: IFI/Plenum, 1973.
156 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

[304] Mark HF, editor. Encyclopedia of chemical technology/Kirk-Othmer. New York: Wiley, 1978.
[305] Balandin VY, Otte D, Bostanjoglo O. J Appl Phys 1995;78(3):2037.
[306] Martynyuk MM. Sov Phys Tech Phys 1994;19:793.
[307] Peterlongo A, Miotello A, Kelly R. Phys Rev E 1994;50:4716±27.
[308] Knacke O, Kubaschewski O, Hesselmann K. Thermochemical properties of inorganic substances.
2nd ed Berlin: Springer Verlag, 1991.
[309] Song KH, Xu X. Appl Phys 1997;A 65:477±85.
[310] Kelly R, Miotello A, Braren B, Gupta A, Casey K. Nucl Instrum and Methods 1992;B 65:187±99.
[311] Kelly R, Miotello A. Appl Phys 1993;B 57:145±58.
[312] Kelly R, Miotello A. Nucl Instrum and Methods 1994;B 91:682±91.
[313] Miotello A, Kelly R. Appl Phys Lett 1995;67:3535±7.
[314] Kelly R, Miotello A. Appl Surf Sci 1996;96-98:205±15.
[315] Kelly R, Dreyfus RW. Nucl Instr Methods 1988;B32:341.
[316] FaÈhler S, Krebs HU. Appl Surf Sci 1996;96:61.
[317] Singh RK, Narayan J. Phys Rev 1990;B 41:3174.
[318] Anisimov VN, Arutyunyan RV, Baranov VY, Bolshov LA, Velikhov EP, Dolgov VA, Ilyin AI,
Kovalevich AM, Kraposhin VS, Malyuta DD, Matveeva LA, Mezhevov VS, Pismennyi VD,
Sebrant AY, Stepanov YY, Stepanova MA. Appl Opt 1984;23:18±25.
[319] Mazhukin V, Smurov I, Flamant G. J Comp Phys 1994;112:78.
[320] Root RG. Modeling of Post-breakdown phenomena. In: Radziemski LJ, Cremers DA, editors.
Laser-induced plasmas and applications. New York and Basel: Marcel Dekker, 1992.
[321] Chichkov BN, Momma C, Nolte S, Alvensleben FV, TuÈnnermann A. Appl Phys 1996;A 63:109±15.
[322] Raizer YP. Soviet Physics JETP 1965;21:1009±17.
[323] Pirri AN. The Physics of Fluids 1973;16:1435±40.
[324] Pirri AN, Root RC, Wu PKS. AIAA Journal 1978;16:1296.
[325] Boni AA. Su FY. Thomas PD. Musal HM. Theoretical study of laser±target interactions. Final
Tech. Report SAI 77-77-567-LJ, Science Application Inc., La Jolla, California, May 1977.
[326] Sedov LI. Similarity and dimensional methods in mechanics. New York: Academic Press, 1959.
[327] Ferriter N, Maiden DE, Winslow AM, Fleck JA. AIAA Journal 1977;5:1597±603.
[328] Poprawe R. Materialabtragung und Plasmaformation im Strahlungsfeld von UV-Lasern. Ph.D
dissertation, Technische Hochschule, Darmstadt, 1984.
[329] Holmes BS, Tarver C, Ehrlich DC, Lindberg HE. The mechanical loads from LSD waves and their
simulation. Final Report F29601-74-C-0051, Stanford Research Institute, March 1976.
[330] Anisimov SI, Khokhlov VA. Instabilities in laser±matter interaction. Boca Raton, London, Tokyo:
CRC Press, 1995.
[331] Schutte K. Prozeûdiagnostik und Technologische Untersuchungen Zur Materialbearbeitung mit
Excimerlasern. Ph.D thesis, UniversitaÈt Erlangen-NuÈrnberg, NuÈrnberg, 1993.
[332] Reilly JP, Ballantyne A, Woodro€e JA. AIAA Journal 1979;17(10):1098.
[333] Ishiguro H, Ohyama K, Nariai H, Teramoto T. J Nucl Sci Technol 1990;27(12):1115±25.
[334] Semak V, Matsunawa A. J Phys D: Appl Phys 1997;30:2541±52.
[335] Kelly R, Rothenberg JE. Nucl Instr Methods 1985;B7/8:755.
[336] Kelly R, Miotello A, Braren B, Otis CE. Appl Phys Lett 1992;60(24):2980.
[337] Bellantone R, Hahn Y. J Appl Phys 1994;78:1436±46.
[338] Poprawe R. Materialabtragung und Plasmaformation im Strahlungsfeld von UV-Lasern. Ph.D
thesis, Technische Hochschule Darmstadt, Darmstadt, 1984.
[339] Luft A, Franz U, Emsermann A, Kaspar J. Appl Phys 1996;A 63:93±101.
[340] Anisimov VN, Baranov VY, Bolshov LA, Ilyin AI, Kopetskii CV, Kraposhin VS, Malyuta DD,
Matveeva LA, Pismennyi VD, Sebrant AY. Phys Chem Mech Surfaces 1985;3(9):2756.
[341] D'Anna E, Leggieri G, Luches A. Thin Solid Films 1992;218:219±30.
[342] Mihailescu IN, Chitica N, Nistor LC, Popescu M, Teodorescu VS, Ursu I, Andrei A, Barborica A,
Luches A, Luisa de Giorgi M, Perrone A, Dubreuil B, Hermann J. J Appl Phys 1993;74:5781.
[343] D'Anna E, Leggieri G, Luches A, Martino M, Drigo AV, Mihailescu IN, Ganatsios S. J Appl Phys
1991;69:1687±96.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 157

[344] Armco Inc. Armco, Massachussetts, USA 1998.


[345] Goodfellow. Materialien fuÈr Forschung und Entwicklung. Cambridge: Goodfellow Metals Ltd.
1988.
[346] Wegst CW. StahlschluÈssel: Key to steel. Marbach: Verlag StahlschluÈssel Wegst, 1989.
[347] Neubauer M. Hyperfein-Untersuchungen an ionenbestrahlten Ag/Fe und In/Fe Schichten. PhD
thesis, UniversitaÈt GoÈttingen, 1996.
[348] Neubauer M, Lieb KP, Schaaf P, Uhrmacher M. Thin Solid Films 1996;275:69±72.
[349] Tesmer J.R., Nastasi M., editors. Handbook of modern ion beam materials analysis. Pittsburgh,
Pennsylvania: Materials Research Society, 1995.
[350] Bolse W. Mater Sci Eng 1994;R 12(2):53±122.
[351] Feldmann LC, Mayer JW. Fundamentals of surface and thin ®lm analysis. New York: Elsevier,
1986.
[352] Bethge K, Rauch F, Misaelidis P. editors. Amsterdam: Elsevier: 1990 (also Nucl Instrum and
Methods B, vol. 50).
[353] Lieb KP. Atomkerne. In: Teilchen, volume 4 of Bergmann-SchaÈfer: Lehrbuch der Experimental-
physik, chapter 2. Walter de Gruyter, Berlin, New York, 1992.
[354] Chu WK, Mayer JW, Nicolet MA. Backscattering spectrometry. Orlando, Florida: Academic Press,
1978.
[355] Lieb KP. Bergmeister F. Osipowicz T. Pampus K. Uhrmacher M. The GoÈttingen ion implanter
IONAS: a versatile tool for nuclear solid state studies. In: Lieb KP, Uhrmacher M, editors, Solid
State Reactions after Ion Implantation Detected by Nuclear Methods, GoÈttingen, 1986. p. 191.
[356] Lieb KP, Bolse W, Corts T, Kehrel A, Uhrmacher M, Weber T. Depth pro®ling of nitride surface
layers by resonant nuclear reaction analysis. In: Broszeit E, et al, editor. Plasma surface engineering,
p. 1055. Oberursel: DGM Informationsgesellschaft, 1989.
[357] Lieb KP. Characterization of solid state reactions in thin ®lms by nuclear methods. In: Stanek J, et
al, editor. Proceedings of the XXIV Zakopane School of Physics 1990, Singapore, 1990. Singapore:
World Scienti®c, 1990. p. 108.
[358] Uhrmacher M, Pampus K, Bergmeister FJ, Purschke D, Lieb KP. Nucl Instr Methods B
1985;9:234.
[359] Rutherford E. Phil Mag 1911;21:669.
[360] Canberra. PIPS particle detectors, Technical speci®cations. 1996.
[361] Rauhala E. Energy Loss. In: Tesmer and Nastasi [349], 1995 (chapter 2).
[362] Bohr N. Mat Fys Medd Dan Vid Selsk 1948;18:8.
[363] Lindhard J, Schar€ M. Mat Fys Medd Dan Vid Selsk 1953;27:15.
[364] Ziegler JF, Biersack JP, Littmark U. The stopping and range of ions in solids. The stopping and
ranges of ions in matter, vol. 1. New York: Pergamon Press, 1985.
[365] Bethe HA. Ann Phys 1930;5:325.
[366] Bethe HA. Z Phys 1932;76:293.
[367] Bethe HA, Heitler W. Proc R Soc London 1934;A146:83.
[368] Bloch F. Ann Phys 1933;16:287.
[369] Bloch F. Z Phys 1933;81:363.
[370] Andersen HH, Ziegler FJ. In: Hydrogen stopping powers and ranges in all elements. The stopping
and ranges of ions in matter, vol. 3. New York: Pergamon Press, 1978.
[371] Ziegler JF. Helium stopping powers and ranges in all elements, The stopping and ranges of ions in
matter, vol. 4. New York: Pergamon Press, 1977.
[372] Ziegler JF. Stopping cross-sections for energetic ions in all elements, The stopping and ranges of
ions in matter, vol. 5. New York: Pergamon Press, 1980.
[373] Ziegler JF, Manoyan JM. Nucl Instrum and Methods 1988;B 35:215.
[374] Bragg WH, Kleeman R. Philos Mag 1905;10:318.
[375] Boutard D, MoÈller W, Scherzer BMU. Phys Rev 1988;B 38(5):2988.
[376] Thwaites DI. Nucl Instrum and Methods 1987;B 27:293.
[377] Ziegler JF, Chu WK, Feng JY. Appl Phys Lett 1975;27:387.
[378] Biersack JP. PC program package SRIM. 1996.
158 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

[379] Vavilov PV. Sov Phys JETP 1957;5:749.


[380] Bird JR, Williams JS. Ion beams for materials analysis. Sydney, New York, Tokyo: Academic
Press, 1989.
[381] Bohr N. Phil Mag 1915;30:581.
[382] Mayer JW, Rimini E. Ion handbook for materials analysis. New York, San Francisco, London:
Academic Press, 1977.
[383] Kumakhov MA, Komarov FF. Energy loss and ion ranges in solids. New York, London, Paris:
Gordon and Breach, 1981.
[384] Payne MG. Phys Rev 1969;185:611.
[385] TschalaÈr C. Nucl Instrum and Methods 1968;61:141.
[386] TschalaÈr C. Nucl Instrum and Methods 1968;64:237.
[387] Chu WK. Phys Rev 1976;A 13:2057.
[388] Bonderup E. Hvelplund Phys Rev 1971;A 4:562.
[389] Borowski M. PraÈzisionsmessungen zweier Resonanzen in den Reaktionen 14 N…p; †15 O und
18
O…p; †15 N. Master's thesis, UniversitaÈt GoÈttingen, 1992.
[390] Doolittle LR. Program package RUMP for RBS analysis, 1986. RUMP 3.4x.
[391] Doolittle LR. Nucl Instrum and Methods 1985;B9:344.
[392] Doolittle LR. Nucl Instrum and Methods 1986;B15:227.
[393] Saarilahti J. Nucl Instrum and Methods 1992;B64: 734. GISA 3.3.
[394] Biersack JP. IBA, PC Programm Package IBA. 1997.
[395] RBX. University of Budapest, PC program package RBX. 1993.
[396] Mayer M. Technical Report Max-Planck-Institut fuÈr Plasmaphysik, IPP 9/113. SIMNRA 4.0. 1997.
[397] Hirvonen JP. Lappalainen R. Nuclear Reaction Analysis: Particle-Gamma Reactions. In: Tesmer
and Nastasi [349], 1995 (chapter 2).
[398] Damjantschitsch H, Weiser M, Heuser G, Kalbitzer S, Mannsperger H. Nucl Instrum and Methods
1983;218:129.
[399] Maurel B, Amsel G. Nucl Instrum and Methods 1983;218:159.
[400] LaMarche PH, Lanford WA, Golub K. Nucl Instrum and Methods 1981;189:533.
[401] Vizkelethy G. Nastasi M. Tesmer JR. Nuclear reaction analysis: particle-particle reactions. In:
Tesmer and Nastasi [349], 1995 (chapter 6).
[402] Amsel G, Lanford WA. Ann Rev Nucl Part Sci 1984;34:435.
[403] Bird JR. Nucl Instrum and Methods 1980;168:85.
[404] Ziegler JF. New uses of ion accelerators. New York: Plenum Press, 1975.
[405] Deconnick G. Introduction to radioanalytical chemistry. Amsterdam: Elsevier, 1978.
[406] Maurel B. Ph.D dissertation, Universite de Paris VII, 1980.
[407] Maurel B, Amsel G, Nadai JP. Nucl Instr Methods 1982;197:1.
[408] Vickridge I, Amsel G. Nucl Instrum and Methods 1990;B 45:6.
[409] Osipowicz T, Lieb KP, BruÈssermann S. Nucl Instr and Methods 1987;B 18:232.
[410] Rolfs C, Rodney WS. ToÈaÈst Nucl Phys 1974;A 235:450±9.
[411] Lamb WE. Phys Rev 1939;55:190.
[412] Kul'ment'ev AI, Storizhko VE, Zabashta OI. Nucl Instrum and Methods 1994;B 85:633±6.
[413] Kittel C. Introduction to solid state physics. New York: John Wiley and Sons, 1971.
[414] Smulders PJM. Nucl Instrum and Methods 1988;B 14:234.
[415] Clarckson RG, Jarmie N. Comput Phys Commun 1971;2:443.
[416] Shulek P, Golovin BM, Kulyukina LA, Medved SV, Pavlovich P. Sov J Nucl Phys 1967;4:400.
[417] Jarmie N, Pindzola MS, Bichsel H. Comput Phys Commun 1977;13:317.
[418] Lewis MB. Nucl Instrum Methods 1981;190:605.
[419] Amsel G, Davies J. Nucl Instrum and Methods 1983;218:177.
[420] Amsel G, David D. Rev de Phys Appl 1969;4:383.
[421] Land DJ, Simmons DG, Brennan JG, Brown MDMeyer O, Linker G, KaÈppeler F, et alIon beam
surface layer analysis 1976;vol. 2:851Plenum PressNew York.
[422] Deconnick G, Oystaeyen BV. Nucl Instrum and Methods 1983;218:165.
[423] Lappalainen R. Phys Rev 1986:B 34.
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 159

[424] Vickridge IC, Trompetter WJ, Brown IWM, Patterson JE. Nucl Instrum and Methods 1995;B
99:454±7.
[425] Rickards J. Nucl Instrum Methods 1991;B 56/57:812.
[426] Vizkelethy G. Nucl Instrum and Methods 1992;B 64:734.
[427] Landry F. Schaaf P. Nucl Instrum and Methods B, 2000 (submitted).
[428] Press WH, Flannery BP, Teukolsky SA, Vetterling WT. Numerical recipes. The art of scienti®c
computing. Cambridge: Cambridge University Press, 1986.
[429] MoÈssbauer RL. Z Physik 1958;151:124±43.
[430] MoÈssbauer RL. Naturwissenschaften 1958;45:538±9.
[431] MoÈssbauer RL. Science 1962;137:731±8.
[432] Wertheim GK. MoÈssbauer e€ect: Principles and Applications. New York and London: Academic
Press, 1964.
[433] Wegener H. Der MoÈssbauer E€ekt und seine Anwendungen in Physik und Chemie. Biblio-
graphisches Institut, HochschultaschenbuÈcher-Verlag, Mannheim, 1966.
[434] May L, editor. An introduction to MoÈssbauer spectroscopy. New York, London: Plenum Press,
1971.
[435] Gonser U, editor. MoÈssbauer spectroscopy, volume 5 of Topics in Applied Physics. Berlin/Heidel-
berg/New York: Springer Verlag, 1975.
[436] GuÈtlich P, Link R, Trautwein A. MoÈssbauer spectroscopy and transition metal chemistry. Berlin:
Springer Verlag, 1978.
[437] Barb D. Grundlagen und Anwendungen der MoÈssbauerspektroskopie. Berlin: Akademie Verlag,
1980.
[438] Gonser U, editor. MoÈssbauer spectroscopy II, Topics in Current Physics, vol. 25. Berlin-Heidelberg-
New York: Springer Verlag, 1978.
[439] Gonser U. MoÈssbauer Spectroscopy. In: Gonser U, editor. Microscopic Methods in Metals. Berlin
Heidelberg New York: Springer Verlag, 1978. p. 409±48 chapter 13.
[440] Gonser U. From a strange e€ect to MoÈssbauer spectroscopy. In Gonser [435], p. 1±51 (chapter 1).
[441] Fries SM, Crummenauer J, Gonser U, Schaaf P, Chien CL. Hyper®ne Interactions 1989;45:301±8.
[442] Blaes N, Fischer H, Gonser U. Nucl Instrum Methods B 1985;9:201±8.
[443] Schatz G, Weidinger A. Nukleare FestkoÈrperphysik. Stuttgart: Teubner StudienbuÈcher, 1978.
[444] Schaaf P. Conversion of MoÈssbauer Hyper®ne Parameters to PAC Hyper®ne Parameters and Vice
Versa. (Unpublished, 1994).
[445] Spijkermann JJ. Conversion electron MoÈssbauer spectroscopy. In: Gruvermann IJ, editor. MoÈss-
bauer e€ect methodology, vol. 7, 1978. p. 85±96.
[446] Krakowski RA, Miller RB. Nucl Instrum Methods 1972;100:93±105.
[447] Tricker MJ. Conversion electron MoÈssbauer spectroscopy and its recent development. In: Stevens
JG, Shenoy GK, editors. MoÈssbauer spectroscopy and its chemical applications. Advances in
Chemistry Series 194. Washington DC: American Chemical Society, 1978. p. 63±100 chapter 3.
[448] Wagner FE. J de Physique 1976;37:C6:673±89.
[449] Gonser U, Schaaf P, Fresenius J. Anal Chem 1991;341:131±5.
[450] Keisch B. Nucl Instrum Methods 1972;104:237±40.
[451] Flinn PA. A MoÈssbauer backscatter spectrometer with full data processing capability. In: Gruver-
man J, Seidel CW, Kieterly DK, editors. MoÈssbauer e€ect methodology, vol. 9. New York: Plenum
Press, 1978. p. 245±58.
[452] Fultz BT. United States patent 4, 393, 306, 1983.
[453] Blaes L, Wagner HG, Gonser U, Welsch J, Sutor J. Hyper®ne Interactions 1986;29:1571±4.
[454] Schaaf P, Blaes L, Welsch J, Jacoby H, Aubertin F, Gonser U. Hyper®ne Interactions
1990;58:2541±6.
[455] Nasu S, Gonser U. J de Physique 1980;41:C8:690±3.
[456] Schaaf P, KraÈmer A, Blaes L, Wagner G, Aubertin F, Gonser U. Nucl Instrum and Methods
1991;B 53:184±6.
[457] Gonser U, Schaaf P, Aubertin F. Hyper®ne Interactions 1991;66:95±100.
[458] Schaaf P, KraÈmer A, Aubertin F, Gonser U. Z Metallkunde 1991;82:815±9.
160 P. Schaaf / Progress in Materials Science 47 (2002) 1±161

[459] Schaaf P, Wenzel T, Schemmerling K, Lieb KP. Hyper®ne Interactions 1994;92:1189±93.


[460] KuÈndig W. Nucl Instrum Methods 1969;75:336±40.
[461] Landry F and Schaaf P. GoÈMOSS Ð PC Fitprogram for the analysis of MoÈssbauer spectra. 1998.
[462] DIN. Deutsches Institut fuÈr Normung, DIN, 1997; 50359-1:1±17.
[463] Behnke HH, Weiler W. MaterialpruÈfung 1988;30:239±42.
[464] Behnke HH. HaÈrterei-Technische Mitteilungen 1993;48:3±10.
[465] Behnke HH. VDI-Berichte, 1995;1194:33±46.
[466] Dengel D. Entwicklung und Stand der UniversalhaÈrtepruÈfung. In Mayr P, VoÈhringer O, editors.
Werksto€kunde, Oberursel: DGM Verlag. 1991. p. 397±409.
[467] Oliver WC, Pharr GM. Journal Materials Research 1992;7:1564±83.
[468] Weiler W, Behncke HH. MaterialpruÈfung 1990;32:303±10.
[469] Weiler W. HaÈrtepruÈfung an Metallen und Kunststo€en. Ehningen: Expert Verlag, 1990.
[470] Weiler W, Fischer H. MaterialpruÈfung 1987;29:308±11.
[471] Sneddon IN. Int J Eng Sci 1965;3:47±57.
[472] Loubet JL, Georges JM, Meille G. Vickers indentation curves of elastoplastic materials. In: Blau
PJ, Lawn BR, editors. Microindentation techniques in Materials Science and Engineeringvol. 889.
Philadelphia: ASTM Special Technical Publication, 1978.
[473] DIN. DIN 4768: RauheitsmeûgroÈûen.
[474] DIN. DIN 4775: PruÈfen der Rauheit an WerkstuÈckober¯aÈchen.
[475] Queitsch R. Master's thesis, UniversitaÈt Erlangen-NuÈrnberg, Erlangen, 1992.
[476] Mele A, Guidoni AG, Kelly R, Miotello A, Orlando S, Teghil R, Flamini C. Nucl Instrum and
Methods 1996;B116:257±61.
[477] Joint Comittee on Powder Di€raction Standards. Selected powder di€raction data for metals and
alloys. American Society for Testing and Materials ASTM, Swartmore, 1978.
[478] Henry NFM, Lansdale K, editors. International tables for X-ray crystallography, vol. 1. Birming-
ham: The Kynoch Press, 1952.
[479] Ibers JA, Hamilton WC, editors. International tables for X-ray crystallography, vol. 4. Birming-
ham: The Kynoch Press, 1974.
[480] Hahn T, editor. International Tables for Crystallography, vol. A. Dordrecht: Reidel, 1985.
[481] Bruker AXS. Di€racPlus with SEARCH, EVA, DQUANT, and PDFMaint. Program package,
1998.
[482] Howard CJ, Hill RJ. A computer program for Rietveld analysis of ®xed wavelength X-ray and
neutron di€raction patterns. Report M112, Australian Atomic Energy Commission, 1986.
[483] Hill RJ, Madsen IC. J Appl Cryst 1986;19:10±18.
[484] Rietveld HM. Acta Cryst 1967;22:151±2.
[485] Rietveld HM. J Appl Cryst 1969;2:65±71.
[486] Akhmanov SA, EmelyÂanov VI, Koroteev NI, Seminogov VN. Sov Phys Usp 1985;28:1084.
[487] Bostanjoglo O, Nick T. J Appl Phys 1996;79:8725.
[488] Brailovsky AB, Gaponov SV, Luchin VI. Appl Phys A: Solids Surf 1995;61:81.
[489] Ang LK, Lau YY, Gilgenbach RM, Spindler HL, Lash JS, Kovaleski SD. J Appl Phys
1998;83:4466.
[490] Chandrasekhar S. Hydrodynamic and hydromagnetic stability. New York: Dover, 1961.
[491] Miotello A, Kelly R, Braren B, Otis CE. Appl Phys Lett 1992;61(23):2784.
[492] Kunze J. Nitrogen and carbon in iron and steel. Berlin: Akademie Verlag, 1990.
[493] DIN. Deutsches Institut fuÈr Normung 1997; 50359:1±17.
[494] Weber T, de Wit L, Saris FW, KoÈniger A, Rauschenbach B, Wolf GK, Krauss S. Mater Sci Eng
1995;A 199:205.
[495] Smurov I, Covelli L, Tagirov K, Aksenov L. J Appl Phys 1992;71:3147.
[496] Schaaf P. et al. The ordering of nitrogen interstitials in the " iron nitride: a MoÈssbauer puzzle. 2000
(in preparation).
[497] Schaaf P, et al. The MoÈssbauer parameters of "-carbonitrides. 2000 (in preparation).
[498] Schaaf P, Landry F. MoÈssbauer investigation of nitriding processes: gas nitriding and laser nitrid-
P. Schaaf / Progress in Materials Science 47 (2002) 1±161 161

ing. In: Miglierini M, Petridis D, editors. MoÈssbauer spectroscopy in materials science, NATO
Science Series High Technologies, 1999. p. 161±72.
[499] Ron M. Iron±carbon and iron±nitrogen systems. In: Cohen RL, editor. Applications of MoÈssbauer
spectroscopy II. New York: Academic Press, 1980. p. 329±92.
[500] DeChristofaro N, Kaplow R. Metall Trans A 1977;8:35±44.
[501] Coey JMD, O'Donnell K, Qinian Q, Touchais E, Jack KH. J Phys: Condens Matter 1994;6:L23±
0L28.
[502] Nozik AJ, Wood JC, Haacke G. Solid State Commun 1969;7:1677.
[503] Clauser MJ. Solid State Commun 1970;8:781.
[504] Figueiredo RS, Drago V. Solid State Commun 1991;80:757±60.
[505] Foct J, Rochegude P, Hendry A. Acta Metall 1988;36:501±5.
[506] Oda K, Umezu K, Ino H. J Phys: Condens Matter 1990;2:10147±58.
[507] Fall I, GeÂnin JMR. Hyper®ne Interactions 1991;69:513±6.
[508] Bauer P, Uwakweh ONC, GeÂnin JMR. Hyper®ne Interactions 1988;41:555±8.
[509] Eickel KE, Pitsch W. Phys stat sol 1970;39:121.
[510] Rochegude P, Billon B. Scr Metall 1986;20:1095.
[511] Chen GM, Jaggl NK, Butt JB, Yeh EB, Schwartz LH. J Phys Chem 1983;87:5326±32.
[512] Rykalin NN, Uglov AA, Galiev AL. Sov J Plasma Phys 1978;4:185±8.
[513] Moin E, Murr LE. Scripta Metall 1978;12:575±6.
[514] Jaschek R, Konrad P, Mayerhofer R, Bergmann HW, Bickel P, Kowalewicz R, Kuttenberger A,
Christiansen J. Proc SPIE 1995;2502:724±31.
[515] Schaaf P, Wiesen S, Gonser U. Acta Metall 1992;40:373±9.
[516] MaÈndl S, GuÈnzel R, Richter E, Moeller W. Surface and Coatings Technology 1998;100-101(1-
3):367±71.
[517] Jervis TR, Williamson DL, Hirvonen JP, Zocco TG. Materials Letters 1990;9:379±83.
[518] Ion JC, Easterling KE, Ashby MF. Acta Metall 1984;32:1949.
[519] Gillner A. Wissenbach K. Kreutz EW. Laser surface melting of cast iron: processing parameters,
structures, hardness. In: Mordike [277], p. 213±21.
[520] Mikhalev MS, Egorova TI. Combustion-, Explosion- and Shock Waves 1975;11:321±4 (Fizika-
Goreniya-i-Vzryva, Trans.).
[521] Watters JL, Wilshaw TR, Tetelman AS. Metall Trans 1970;A 1:2849±55.
[522] Hintz G, Tkotz R, Keusch C, Negendanck M, Staudigel J, Christiansen J, Ho€mann DHH, Eisner
K, Lang A, Schutte K, Bergmann HW. Proc SPIE 1997;3092:169±72.

You might also like