You are on page 1of 13

Materials Science and Engineering A286 (2000) 269 281 www.elsevier.

com/locate/msea

A study of internal hydrogen embrittlement of steels


G.P. Tiwari a,* , A. Bose b, J.K. Chakravartty a, S.L. Wadekar a, M.K. Totlani b, R.N. Arya a, R.K. Fotedar b
a

Metal Physics Section, Materials Science Di6ision, Bhabha Atomic Research Centre, Trombay, Mumbai 400 085, India b Materials Processing Di6ision, Bhabha Atomic Research Centre, Trombay, Mumbai 400 085, India Received 22 October 1999; received in revised form 2 February 2000

Abstract A novel procedure for hydrogen charging and studying the Internal Hydrogen Embrittlement (IHE) of steels is described here. A cylindrical notched tensile sample with an extended end is employed for hydrogen charging. The extended portion of the sample forms the cathode in an alkaline bath and a constant uni-axial tensile load is applied during hydrogen charging. The stress gradient set up by the notch, which is not in contact with the electrolyte, enhances the hydrogen concentration at various trapping sites of the matrix beyond the solubility limit. Subsequent to charging, the specimen is kept under the same load as that during charging, for another 24 h to stabilize the population of hydrogen within the specimen matrix. At the end of this stage, the specimen is tensile tested to failure at room temperature. Two different steels namely maraging and mild steels have been chosen to study the effect of hydrogen ingress on mechanical properties. While an increase in tangent modulus (linear portion of the stressstrain diagram), yield strength, work hardening rate and ultimate tensile stress (UTS) has been observed on hydrogenation, a decrease in total elongation has been noticed for both the steels studied. Fractographic investigation has revealed that the fracture mode is predominantly ductile dimple (failure by micro-void coalescence) in both the materials and that the hydrogen reduces the size of the dimples. The observations of this investigation are signicant in two respects: rstly, it demonstrates the efcacy of a hydrogen charging method for steels which can introduce hydrogen to a level much higher than its solubility limit and secondly, it reports for the rst time enhancement of modulus and work hardening by hydrogen charging. These observations have been rationalized on the basis of current understanding on the effect of hydrogen on plastic properties and hypothesis of the models of IHE. It is suggested that the trapping of hydrogen by dislocations and other structural features of the matrix and the mutual interactions of their strain elds can account for the observed effects on yield strength, tangent modulus, work hardening rate, UTS and ductility. 2000 Elsevier Science S.A. All rights reserved.
Keywords: Internal hydrogen embrittlement; Steel; Trapping; Mechanical properties

1. Introduction The effects of hydrogen on the mechanical properties of iron and steels have been widely investigated [13]. As regards the embrittling effect of hydrogen it is well known that the chief effects are a decrease in ductility and true stress at fracture. Hydrogen embrittlement (HE) of steels can be classied into three main types [4,5]. 1. Hydrogen reaction embrittlement arises because of the generation of hydrogen on the surface as a result
* Corresponding author. Tel.: +91-22-5505050; fax: + 91-225505151. E-mail address: gptiwari@magnum.barc.ernet.in (G.P. Tiwari)

of a chemical reaction [6,7]. The resulting hydrogen can either form blisters in the sub-surface region or gaseous methane in the interior. Precipitation of hydrogen as hydride in hydride forming elements such as zirconium and titanium [8,9] are other examples where chemical reaction aids the hydrogeninduced embrittlement of the matrix. 2. Environmental embrittlement takes place in the hydrogen containing atmospheres through adsorption of molecular hydrogen on the surface and its absorption within the lattice after dissociation into atomic form [5]. 3. Internal hydrogen embrittlement (IHE), in contrast takes place in the absence of a hydrogenated atmosphere and is brought about by hydrogen which has

0921-5093/00/$ - see front matter 2000 Elsevier Science S.A. All rights reserved. PII: S 0 9 2 1 - 5 0 9 3 ( 0 0 ) 0 0 7 9 3 - 0

270

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

entered the lattice during processing or fabrication of steel [10]. Having entered the lattice, hydrogen embrittles the steel over a period of time which is a function of concentration, temperature and state of stress within the matrix. The details of the effect of hydrogen on mechanical properties and the mechanism of degradation are reaTable 1 Chemical composition of the steels used Element Composition (wt.%) Mild steel C Si Mn S P Ni Mo Co Ti Al Fe 0.12 0.25 0.40 0.04 0.04 Balance Maraging steel grade 350 0.005 18.39 3.99 12.32 1.63 0.12 Balance

sonably well understood in the case of hydrogen reaction embrittlement. However, generally accepted mechanisms for the other two classications of HE, have not yet been established because of the contradictory experimental results obtained by various researchers [2,3,5]. This is more so in the case of IHE and one of the main reasons for the discrepancies in the reported results has been attributed to the possible structural damage caused by hydrogen charging which masks the intrinsic effect of hydrogen on mechanical properties [11,12]. This paper presents the results of experiments designed to elucidate the intrinsic effect of hydrogen on the mechanical properties of ferritic steels by employing a new technique of hydrogen charging. Two different steels namely mild steel and maraging steel with widely different strength levels have been chosen for this study.

2. Experimental procedure As received materials in tempered condition have been employed in this study. The nominal compositions of these two steels are given in Table 1. The three essential steps in the experimental procedure in this study on internal hydrogen embrittlement are: 1. The electrolytic charging of specimen with hydrogen under uniaxial loading conditions. 2. After charging hydrogen for a specied number of hours, the charging cell is removed and the specimen is held in a xture, shown in Fig. 1, under a predetermined load for a period of 24 h. 3. Finally the specimen is removed from the charging set-up and immediately strained up to fracture at a nominal strain rate of 3 10 5 s 1 in an Universal Testing Machine. Fig. 2 depicts the geometry as well as the dimensions of the specimen. One end of the specimen is extended. The hydrogen charging assembly is wall mounted as shown in Fig. 1. The extended portion of the tensile specimen forms the cathode of the electrolytic cell used for hydrogen charging for various periods of time. The charging time was selected in such way that it was well in excess of the estimated time required for hydrogen to reach the notched section of the loaded tensile samples. An electrolytic solution of 1 wt.% NaOH with 1 g l 1 As2O3 as recombination poison was used. A current density of 1.510 2 A cm 2 was employed for hydrogen charging. The specimen was held rigidly by its threaded sections. This arrangement prevented any lateral movement and also helped in maintaining the proper vertical alignment of the specimen. A dead weight was applied below the elastic limit of the material. This load was arrived at from the knowledge of the

Fig. 1. Schematic of xtures used for hydrogenation and loading mechanism.

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

271

neous. The main difference in the method of hydrogen charging employed by GE with that used in this study lies in the fact that the test portion of the sample in our experiments is fed with hydrogen through diffusion. Secondly, holding the specimen under load (for a period of 24 h) subsequent to hydrogen charging, allows sufcient time for the hydrogen population to stabilise itself within the matrix. Hence, the distribution of hydrogen in the matrix is expected to undergo relatively small changes during the tensile testing, whereas in GEs experiments, the concentration as well as the distribution of hydrogen remains in a state of ux till the end of experiments. An optical microscope was used to characterize the microstructures of the starting material as well as of the hydrogenated material. Fractography of the deformed specimens was carried out using a scanning electron microscope (SEM).

2.1. Estimation of hydrogen diffusi6ity and concentration


Estimation of hydrogen diffusivity in the mild steel stock used in the present investigation was carried out using the method suggested by Devanathan and Stachurski [16]. In this technique, a specimen having rectangular cross-section (approximately 1 1 cm) and 1-cm thick is cathodically charged with hydrogen up to the saturation limit. This saturation limit is reached only after hydrogen has occupied all the available sites on reversible as well as irreversible traps and the concentration in the lattice has risen up to the solubility limit. Fig. 3(a) shows the current versus time plot for the cathodic charging of the 1-cm-thick mild steel specimen. The electrolyte used was the same as that used for hydrogen charging. Open circuit potential was 235 mV (w.r.t. to SCE) and the charging was performed at 400 mV (w.r.t. SCE). The concentration of diffusible hydrogen in the sample controls the charging current and saturation is reached around 28 mA. This state is equivalent to steady state diffusion so that all the hydrogen entering the specimen goes at the other end. According to Deluccia and Berman [17], the diffusivity, D of hydrogen in steel is related to tmax, the time required to reach the saturation level, as D= L 2/(4tmax) (1)

Fig. 2. Schematic showing details of the specimen used for charging and testing.

elastic limit of the material and stress concentration factor for the notch used. The presence of a notch in the centre creates a stress gradient within the specimen. The tendency of the hydrogen to migrate up the stress gradient facilitates hydrogen charging of the specimen [3,13,14]. The procedure outlined above thus closely simulates IHE. Keeping the bare (uncoated) specimen under load in the absence of the source of hydrogen will reduce out-gassing of hydrogen and allow the hydrogen to be present in the lattice in excess of the solubility limit. At the end of this period, most of the hydrogen present in the specimen in excess of solubility is expected to diffuse out of the ferritic lattice and residual hydrogen exists in strongly bonded trapping sites of the steel. Hence the embrittlement effects reported here are primarily due to the presence of hydrogen in tightly bonded irreversible traps. The procedure used for hydrogen charging employed here is somewhat similar to that used by Groeneveld and Elsea [15]. These authors (GE) used a specimen which was extended at both ends and held under tensile load while charging. The charging was done electrolytically as is the case here. However, the charging was done in the middle section of the specimen which was used for the test. The specimen was held under load in the electrolytic bath to determine the cracking time as a function of load and charging conditions. Thus, in their investigation, the charging of hydrogen into the specimen and its trapping within the matrix were simulta-

where L is the thickness of the specimen. With the use of data in Fig. 3, D is estimated to be 310 4 cm2 s 1, on the basis of tmax = 15 min. This diffusivity value is in very good agreement with an earlier reported value [18]. In order to validate this estimation, an anodic oxidation (dehydrogenation) curve of the same specimen was also carried out as shown in Fig. 3(b). Prior to oxidation, a thin oxide layer was grown on both faces of the

272

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

specimen to minimize the interfacial reaction which may otherwise inuence the rate of oxidation. Note that initial current is nearly the same as the saturation current in Fig. 3a. The rate of anodic reaction is diffusion-controlled and hence the time required for saturation and complete dehydrogenation is nearly the same. The concentration of hydrogen in the sample at any time t can be calculated from the anodic curve of Fig. 3b and the use of the following formula again from Deluccia et al. [17] J/nF= Co{D/(y tmax)}1/2 (2)

steel [19]. This is the concentration of diffusible hydrogen in the sample. The estimations of diffusivity and average hydrogen content have helped to establish the efcacy of the charging method for introduction of hydrogen into the steel specimens. The ux equation for the diffusion of hydrogen under combined action of concentration and stress gradient can be written [20,21] as Jc = (DC/RT) [RT(d ln C/dx)V*/3 (d|/dx)] (3) where, Jc is the hydrogen ux, D is the diffusivity of hydrogen in steel, C is the concentration of hydrogen at any point X, V* is the activation volume for diffusion of hydrogen in steel, | is the tensile stress at any point and R and T have their usual meaning. Differentiation of Eq. (3) yields dC/dt =D[(d2C/dx 2)+ K(dC/dx)] (4)

with J, anodic current; n, no. of electrons, Co, concentration of hydrogen and F, Faradays constant. The value of C0 is calculated to be 0.9 ppm which is nearly the same as the solubility limit of hydrogen in

Fig. 3. (a) A plot of current versus time during cathodic charging of mild steel specimen; (b) Anodic oxidation (dehydrogenation) curve for mild steel.

where, K= (V*/3RT) (d|/dx) If d|/dx is constant then K is also constant and the solution of the second-order non-linear differential Eq. (4) can be found analytically [20] or can be solved by appropriate numerical methods. A solution of this equation with proper boundary conditions can then be utilised to estimate the hydrogen concentration at any section of the specimen. However, this has not been done for two reasons. Firstly, the cross-section of the tensile specimen varies from place to place and hence Eq. (4) cannot be applied directly to the present case. Secondly, as mentioned earlier, the specimen is held under load without the presence of electrolytic charging cell. During this period, most of the diffusible hydrogen will escape away from the specimen in the absence of the concentration gradient. Tensile loading preferentially activates the trapping sites whose axes are at right angles to the load [2224]. It is believed that all the hydrogen present in the sample is held in such traps for which Eq. (4) is not applicable in its present form. However, it is clear from these expressions that there will be a considerable enrichment of hydrogen at the notched section of the specimen. In order to obtain an estimate of the hydrogen content in the bulk at the time of fracture, vacuum fusion analysis has been carried out on the samples taken from regions just below the fractured notched portion of the specimen. An average value of 5 ppm hydrogen has been obtained in such samples and this value is much higher than the solubility limit of hydrogen in ferrite matrix at ambient conditions ($1 ppm). This value may be considered as the lower bound solubility limit of hydrogen at irreversible trapping sites because out-gassing of hydrogen after fracture and the actual hydrogen concentration will be much higher within the loaded sample before the commencement of the tensile test.

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

273

ally, such tests also yield information on the deformation behaviour of the material in the absence of hydrogen. A dead weight of 4300 N (corresponding to approximately 65% of the yield load) was chosen for hydrogen charging of the mild steel and this load set up stresses of 220 and 150 MPa at the notch and in the region away from the notch, respectively. The corresponding stress values for maraging steel samples were 250 and 175 MPa and were generated using a dead weight of 5000 N which was approximately 65% of the yield stress. An inspection of Figs. 4 and 5 reveals that the hydrogen inuences the deformation behaviour of mild steel and the maraging steels in the following way.

3.1. Effecti6e modulus


The initial parts of stressstrain curves in Figs. 4 and 5 represent the elastic deformation and the slope yields the value of the tangent modulus (taken here as the slope of the initial linear portion of the stressstrain curve). It is clear from Tables 2 and 3, hydrogen enhances the magnitude of the tangent or effective modulus in both mild steel and maraging steel. In mild steel, for charging times up to 13 h, there was no noticeable change in the values of the effective modulus. However, a signicant increase in tangent modulus was observed for longer charging times (e.g. for 22 and 40 h). The increase in the tangent modulus is higher for maraging steel compared to that in mild steel for similar charging times. In mild steel, the increases are 33 and 80%, respectively for 22 and 40 h of charging and the corresponding increase in maraging steel for 22 h of charging is about 61%. To our knowledge, this observation of the enhancement in tangent modulus of steel due to the presence of hydrogen during tensile deformation is new and has not been reported before. The slope of the linear portion of the stress strain plot is dependent on both Youngs modulus and the micro-yielding process. In a polycrystalline material (as in the case of these two steels), the early stage of deformation (micro-yielding) is associated with yielding in the isolated individual grains which are most favourably oriented with respect to the direction of stressing and movement of dislocations in those grains from the source to slip barriers such as grain boundaries, precipitates, etc. At these barriers, the piled up dislocations create stress concentrations to activate slip in the adjacent grain. The process repeats till all the grains in the load bearing section of the specimen yield at the macroscopic yield stress.

Fig. 4. Stress strain plot for notched mild steel specimens charged with hydrogen for different durations and an un-notched specimen without introduction of hydrogen.

Fig. 5. Stress strain plot at room temperature for notched maraging steel samples with and without hydrogen.

3. Results The results of tensile tests on mild steel and maraging steel are given in Figs. 4 and 5. Fig. 4 shows the tensile test results for mild steel with and without notch. For maraging steel, the tests were performed only on notched specimens with and without hydrogen. The purpose of such tests was to estimate the limiting elastic load to be applied during hydrogen charging. Addition-

3.2. Loss of ductility and enhancement of strength


The most signicant effect of the presence of hydrogen on the deformation behaviour is the loss in the

274

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

ductility (or notch ductility) which decreases from a value of 11.2%, for the notched and uncharged specimen of mild steel to 5.7, 5.5 and 4.1% for 13, 22 and 40 h of hydrogen charging, respectively. In percentile terms, this amounts to a decrease of 49 64%. The loss in ductility in maraging steel is even greater and amounts to nearly 69% for 22 h of charging compared to uncharged material. The detailed analyses of tensile test data of Figs. 4 and 5 are given in Tables 2 and 3. It is observed that hydrogen charging causes a large increase in the yield strength (160 and 75% increase in mild steel and maraging steel, respectively) compared to uncharged condition. With increasing charging times, however the yield strength remains virtually unchanged in mild steel. In contrast to the effect of hydrogen on ductility, the UTS (Notched Tensile Strength) values of the hydrogen charged specimen show a rising trend with increasing charging times for both mild steel and maraging steel as found in the case of effective modulus and yield strength. However, the increases in UTS in both these materials are not as large as in the case of effective modulus. It must be mentioned here that an increase in UTS should not be taken as an indication of improvement of toughness also. Since toughness of a notched specimen depends strongly on the notch ductility, a reduction of more than 50% in this parameter as observed here at longer charging times (22 h and higher) indicates a strong embrittling effect of hydrogen.

stressstrain curve, is enhanced signicantly in the presence of hydrogen. For mild steel, the work hardening rate is found to increase from 5.6 103 MPa for un-hydrogenated state to 5.8103, 6.2 103 and 1.2 104 MPa for 13, 22 and 40 h of hydrogen charging, respectively. The corresponding percentile increases are 4, 11 and 116% compared to un-hydrogenated mild steel material. For maraging steel, after 22 h of charging, the work hardening rate jumps to 1.3104 MPa from the uncharged value of 5.9103 MPa which amounts to an increase of more than 100%.

3.4. Fracture mode


Results of fractographic investigations are shown in Figs. 6 and 7 for mild steel and maraging steel, respectively. It is observed that fracture modes of both unnotched (Fig. 6(a)) and notched (Fig. 6(b)) specimens in uncharged conditions and those of hydrogen charged for 13 h (Fig. 6(c)) are essentially ductile and fracture occurs by micro-void coalescence (mvc). At longer charging time (22 h), the fracture mode remains predominantly ductile in nature with some quasi cleaved/ cleaved area (Fig. 6(d)). Although no quantitative measurements have been made, it is apparent that the dimple size is larger in the uncharged samples compared to that in the charged samples. In addition to this, dimple size was found to decrease with increasing charging time. A similar general trend in fracture mode of maraging steel has also been observed as shown in Fig. 7(a,b). It is noted that the failure in this material is also by micro-void coalescence and the presence of hydrogen decreases the dimple size. However, it must be mentioned that damage such as cracking (or de-cohesion) at the particlematrix interface and grain boundaries have not been seen even at longer charging

3.3. Enhancement of work hardening rate


Tables 2 and 3 show yet another new and interesting result on the effect of hydrogen on the work hardening rate. The work hardening rate, which is the average work hardening rate in the plastic portion of the
Table 2 Effect of hydrogen on mechanical properties of mild steel Specimen Effective modulus (GPa) 11 11.5 11.5 15.3 20.6

Work hardening rate (MPa, 103) 3.1 5.6 5.8 6.2 12

UTS (MPa) 652 845 880 905 935

Total elongation (%) 13.3 11.2 5.7 5.5 4.1

Without notch and without hydrogen Notched and without hydrogen Notched and 13 h hydrogen Notched and 22 h hydrogen Notched and 40 h hydrogen

Table 3 Effect of hydrogen on mechanical properties of maraging steel Specimen Effective modulus (GPa) 13.2 21.2 Work hardening rate (MPa, 103) 5.9 13.0 UTS (MPa) 1247 1340 Total elongation (%) 19.3 6.1

Notch without hydrogen Notch and 24 h hydrogen

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

275

Fig. 6. Scanning electron micrograph of the surface of mild steel samples pulled to failure in tension at room temperature; (a) without notch and without hydrogen; (b) with notch and without hydrogen; (c) with notch and with hydrogen (13 h charging) and (d) with notch and with hydrogen (22 h charging).

times. The results of fractography of notched samples presented here, correspond to the areas near the notch tip and the fracture at region away from the notch (i.e. near the centre of the specimen) occurred essentially by mvc. The effects of hydrogen on fracture mode and on the size of the dimples are indicative of the fact that hydrogen inuences the local plasticity to a considerable extent. The results of fractographic analysis substantiates the observed effects of hydrogen on ductility given in Tables 2 and 3. A reduction in the dimple size with associated loss of ductility on hydrogen charging has been reported earlier in the case of IHE in various materials by several authors [25,26].

4. Discussion This investigation suggests a new method of charging hydrogen and studying the intrinsic effect of hydrogen on mechanical properties of steel as during IHE. Hydrogen is charged into the test specimen by stress assisted diffusion of hydrogen generated remotely by electrolytic charging. Hydrogen charged in this manner is not expected to cause any damage to the microstructure as the test piece is not in direct contact with the electrolyte and is thus not exposed to high fugacity hydrogen that prevails during cathodic charging. In IHE, it is believed that hydrogen having once entered the lattice remains dormant till the ambient conditions allow its transport (by diffusion) to the stressed region.

From mechanistic considerations, the two processes which control the inuence of hydrogen on mechanical properties are its rate of diffusion and its interaction with the trapping sites within the matrix [25,27,28]. The design of the present experiment is such that it allows to some extent to separate these two processes. The presence of stress accentuates the rate of diffusion and the extended charging periods allow sufcient time for trapping events to occur. Subsequently, the holding of the specimen under load without the presence of the source of hydrogen, stabilises the population of hydrogen atoms within the matrix. Hence, the diffusion of hydrogen, per se, has little effect on the deformation characteristics of hydrogen charged steels reported here. The change in mechanical properties is mainly due to the changes in the properties of the structural features such as dislocation, particlematrix interface, grain boundaries and magnetic domain walls brought about by their association with hydrogen.

4.1. New results on IHE and hydrogen trapping 4.1.1. Summary of the results The present investigation has found that for the steels studied, the intrinsic effect of hydrogen is to increase the tangent modulus, strength and work hardening rate and to decrease the elongation. The results also strongly suggest that hydrogen causes hardening in steels. Such hardening effects in ferritic steels by hydrogen have been reported by many workers [2,3,29,30].

276

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

The effects of hydrogen on ductility and work hardening are also consistent with earlier reported research [3,11]. However, the observations made on the effect of hydrogen on the tangent modulus and UTS are novel and signicant and have not been reported before. The increase in UTS as found in this study is unusual and is apparently at variance with the generally observed effect of hydrogen which is known to reduce the fracture stress in ferritic steels [1 3,12,14]. The most interesting observation of this study is the signicant increase of the tangent modulus of both the steels on hydrogenation. This is opposite to the observation made by Bastein and Azou [31] who reported that Youngs modulus was unaffected by hydrogenation. As regards the mode of fracture, it is found that hydrogen does not alter the fracture mode which is predominantly transgranular ductile failure. However, a hydrogen-induced reduction in dimple size is noted and the origin of these dimples is invariably the particle matrix interface rather than the grain boundaries as no intergranular failure has been observed. The participation of smaller carbide/precipitate particles as void initiation sites appears to be responsible for smaller dimple size and

lower ductility in the case of hydrogenated materials [5,25].

4.1.2. Hydrogen trapping in steels The summary of the results obtained here suggests that the reason for the unusual effects of hydrogen stems from the accumulation of a high concentration of hydrogen within the matrix through the new method of hydrogen charging. It is known that a large amount of hydrogen (much higher than the equilibrium concentration) can be retained at ambient temperature in ferritic matrix because of the presence of numerous structural defects which trap hydrogen[14,32,33]. Pressuoyre has classied trapping sites in terms of their physical nature in iron and steels and suggested that vacancy, alloying elements, dislocations, micro-voids and interfaces are possible trapping sites [32]. In steels, the major trapping sites are found to be grain boundaries, sub-grain boundaries, iron carbide and alloy carbides [32]. It has been demonstrated by Podgurski and Oriani [33] that even structural features like insoluble impurities can sustain hydrogen concentration far beyond the solubility limit. As regards the effect of stress, in addition to enhancing local solubility at a trapping site, it may modify trapping characteristics of a structural feature. For example, the grain boundary, in general, has a much higher trapping efciency compared to an interface in steels. However, if the grain boundaries are unfavourably oriented with respect to the tensile stress direction, the interfaces may act as more efcient traps for hydrogen [22]. From the preceding section, it is clear these factors are expected to increase the net amount of trapped hydrogen in steels containing a high density of structural defects as used in this investigation and that we are dealing with samples having signicantly higher amounts of hydrogen (much greater than 5 ppm). Further, it is logical also to infer that a high density of trapped hydrogen cannot exist in isolation and hence there must be some hydrogen present in the lattice as well. A dynamic equilibrium of the concentration of hydrogen should then exist between the lattice and trapping sites. The ratio of concentration level at these sites will depend upon the relative binding energy for a proton. 4.1.3. Effect of trapped hydrogen on mechanical properties The present results are most satisfactorily interpreted if hydrogen or hydrogensolute complex interacts with dislocations to immobilize them and interferes with the nucleation of dislocations from their possible sources. The hydrogen in the lattice, apart from interacting with dislocations by creating dilatational eld [3], can modify the characteristics of dislocations in a variety of ways. Earlier studies on internal friction of hydro-

Fig. 7. Scanning electron micrograph of the surface of maraging steel samples pulled to failure in tension at room temperature; (a) notched samples without hydrogen and (b) notched samples with hydrogen (22 h charging).

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

277

genated iron and steels suggest that dislocations are trapping sites for hydrogen and hydrogen atoms are very effective in anchoring them at ambient temperatures [3436]. The trapped hydrogen induces elastic dipoles and mutual interactions of dipoles centred on dislocations and other structural defects increases the level of back stresses in the lattice. Ke [37] has shown that some hydrogen lodged at dislocations resides in molecular form which is also expected to reduce their mobility. Kornmuller, from magnetic relaxation studies on the interaction between hydrogen and magnetic domain walls concluded that the presence of hydrogen in domain walls enhanced the strength of the elastic stress eld experienced by neighbouring dislocations [38]. In view of the factors mentioned above, the increase in tangent modulus as observed on hydrogenation is obvious as hydrogen is likely to impede glide of individual dislocations during the micro-yield and also prevent generation or unlocking of dislocations from its source in the adjacent grains. In transition metals like niobium, vanadium and titanium, hydrogen in solution has been found to increase as well as decrease Youngs modulus [39]. We admit that our experimental result do not produce any evidence to include explicitly the contribution of hydrogen to enhance Youngs modulus. However, it is apparent from Figs. 4 and 5 that the tangent modulus (the slope of the initial portion of stressstrain diagram) which is dependent on both Youngs modulus and micro-yielding has increased. It is pertinent to note that Laurent et al. [40] observed a , very large concentration (about 600 atoms per A of dislocation) of trapped hydrogen in the interface and dislocations of hydrogenated Armco iron by using high resolution electron autoradiography. A much larger concentration of hydrogen than this is expected in such traps in stressed condition as in this investigation. The state of the lattice surrounding these traps and their possible effect on mechanical properties is not yet known. With regard to the effect of hydrogen on yield strength, it was suggested by Oriani et al. [41] that solute hydrogen can increase the yield or ow stress of steels by hydrogen drag upon moving dislocation or impeding cross-slip. The hardening is also expected from the fact that trapped hydrogen may make dislocation nucleation difcult from grain boundary dislocations and other possible sources. Further it is known that multiple slip will be favoured in a structure containing second phase/carbide particles because of elastic and plastic incompatibility effects. Hirth [3], from a review of a large number of studies, has inferred that hardening occurs when deformation is by multiple slip (at relatively large strain) and that when work hardening is large and the dislocations form a complex array. It is known that a relatively large strain prevails ahead

of a notch at general yield and the yield stress for the notched sample reported in this work thus corresponds to a strain of approximately 5% (estimated from notch geometry used). Therefore, in the investigated steels containing various carbides and precipitates, hardening by hydrogen is expected. The increased work hardening rate in the hydrogenated material relative to the hydrogen-free material can be explained by noting that hydrogen immobilizes dislocations and adversely affects the recovery processes by inhibiting cross slip. As a consequence, there will be build up of a high density of immobile dislocations in the form of loops, tangles, piled up dislocations at carbide/second phase particle etc. which are going to impede glide dislocations. With increasing deformation, the mean free path of dislocations in the matrix will decrease progressively and increasingly higher force will be required to drive dislocations through these barriers. AS UTS is controlled by both yield strength and work hardening, an increase in UTS on hydrogenation is thus apparent. The observed increase in UTS, however, in both the steels by hydrogen is not as effective as in the case of yield strength. This observation is common in any strengthening mechanism (e.g. dynamic strain ageing, solid solution strengthening, etc.) where a relative change in yield strength is usually larger than that in UTS and a consequence of this strengthening is loss of ductility. Similar observations have been made by West et al. [27] on the effect hydrogen on strength and ductility in 21-6-9 stainless steels. It has been pointed out earlier that an increase in UTS should not be taken as an indicator of increase in toughness on hydrogenation. The loss of ductility or embrittlement by hydrogen results from the attainment of critical stress for tensile instability at lower strain compared to hydrogen free material. Similar observations were also made in an earlier study by West and Louthan [27]. Although the steels selected for this study are very different in terms of their chemical composition and mechanical properties, the changes in mechanical properties brought about by hydrogen are strikingly similar. However, the effect of hydrogen in enhancing strength and modulus and decreasing ductility are greater in maraging steel compared to that in mild steel for the same charging time. This is expected as maraging steel is capable of dissolving large amounts of hydrogen by trapping because of the presence of a number of alloying elements besides titanium which have a strong afnity for hydrogen [42,43]. Further the maraging steel has a much ner microstructure and contains a larger volume fraction of second phase particles with a much smaller inter-particle spacing than exists in the mild steel of this study. Therefore a larger increase in strength and work hardening rate is expected in maraging steel on hydrogenation.

278

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

We also note that the effect of hydrogen on deformation characteristics is augmented directly in proportion to its concentration on dislocations and other defects demonstrating that the observed effects of hydrogen are real and intrinsic. The relatively large changes in strength, tangent modulus, work hardening and ductility are related to the high levels of hydrogen trapped in the matrix by the conditions of the experiment. It may be useful at this stage to identify the microstructural features that control the plastic properties and fracture in hydrogenated steels. It is known that the binding energy of hydrogen atom to the dislocation is smaller than particle/matrix interfaces and these interfaces in turn provide traps or hydrogen which are less efcient than grain boundaries. This arises primarily from the fact that the grain boundary has a much higher energy than the interfacial energy of the particle/ matrix interface and thus provides a deeper trap. However, the observations that fracture occurred after considerable macroscopic plastic deformation and the absence of grain boundary fracture in hydrogen charged steels in this investigation suggest that the stress concentration produced by dislocations during deformation are not sufcient to initiate a crack at grain boundaries. Further, the particle/matrix interface or dislocations acting alone cannot explain the mode of fracture on hydrogenation. Therefore, it is reasonable to conclude that the interaction of dislocations or dislocation arrays with carbide/precipitate particle controls plastic deformation and builds up stress concentrations locally to initiate fracture in these steels. The efcacy of the stressed interface in trapping hydrogen has been demonstrated by Ohma et al. [22].

theory of lattice de-cohesion by hydrogen. These models basically propose that dissolved hydrogen in the lattice can reduce the stress needed to initiate or propagate a crack by lowering the cohesive strength of atoms. Gerberich and coworkers [49] examined in detail the role of the state of stress on hydrogen-induced cracking (HIC) in the region of a crack tip. In order to quantify the effect of state of stress, they modeled the crack tip behaviour for HIC by considering small-scale yielding and discretised dislocation computer simulation of stress eld. They further hypothesized that the most severe micro mechanical conditions for HIC prevailed under plane strain condition with mode I loading which was in accordance with TroianoOrianis de-cohesion model.

4.2.3. Reduction of surface energy adsorption Adsorption of hydrogen on the surface reduces the surface energy, hence the total energy required for the fracture is lowered [50]. Just as in the case of the precipitation mechanism mentioned above, this mechanism also requires the presence of micro-porosity. 4.2.4. Association of hydrogen with dislocations Bastien and Azou [51] suggested that embrittlement can result from segregation of hydrogen forming Cotterell atmosphere around dislocations during plastic deformation. This segregated hydrogen atom in a cavity can lead to the formation of a void containing molecular hydrogen gas which under pressure would cause the matrix to rupture. Based on this concept, Tien et al. [52] developed the dislocationtransport model of hydrogen embrittlement. An enrichment of hydrogen in plastically deformed region is predicted when dislocations lose their hydrogen atmosphere during straining. A slightly different model was proposed by Louthan et al. [53] who suggested that a high concentration of hydrogen might result from dislocationhydrogen interaction which could modify the processes of both plastic deformation and fracture. In contrast, Beachem [54] proposed that the basic hydrogendislocation interaction in steels is one of easing motion of dislocations or generation of glide dislocations or both thereby inducing lattice softening. An accumulation of sufcient concentration of hydrogen in the lattice just ahead of a crack tip is assumed to aid the prevailing deformation processes [54]. By noting similarities among hydrogen assisted cracking (HAC), stress corrosion cracking and adsorption induced liquid metal embrittlement for various alloys, Lynch [55] has suggested that HAC involves adsorption of hydrogen at the crack tip. In Lynchs model adsorbed hydrogen is expected to weaken the inter-atomic bonds thereby facilitating dislocation nucleation at the crack tip. This model is somewhat similar to Beachems but differs in choosing the location of the effect of hydrogen. Further evidence of

4.2. Mechanisms of hydrogen embrittlement


To date various models have been put forward to explain the mechanisms of hydrogen embrittlement; however no single theory can fully explain the mechanism of degradation of mechanical properties of steels by hydrogen [44,45]. In the following we take up a few relevant mechanisms to examine the possible reasons for embrittlement.

4.2.1. Precipitation of hydrogen at internal defects When the hydrogen atoms distributed in the lattice precipitate as gaseous hydrogen into pre-existing micropores/voids, the pressure exerted by the gas adds to the externally applied tensile load on the specimen and consequently the effective fracture stress is reduced [46]. 4.2.2. De-cohesion model Troiano [47] suggested that hydrogen dissolved in the regions of high stress in a lattice could reduce the strength of inter-atomic bonds. Oriani [48] further developed Troianos concept and suggested a mechanistic

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

279

hydrogendislocation interactions has been provided by Birnbaum and coworkers [56 58] by carrying out in situ deformation and fracture in TEM environmental cell containing hydrogen gas. These results together with the measurements of thermally activated deformation parameters associated with dislocation motion, following the trend of macroscopic stress strain behaviour and theoretical treatment of hydrogen and dislocation interactions prompted these researchers to suggest a hypothesis for rationalising the hardening softening behaviour on hydrogenation. According to this hypothesis, the presence of hydrogen in solid solution decreases the barriers to dislocation motion thereby increasing the amount of deformation that occurs in a localized region adjacent to the crack surface. This hydrogen enhanced localized plasticity (HELP) has been attributed to the presence of a hydrogen atmosphere around dislocations which shields its interaction with other elastic centres such as dislocations, solute atoms, etc. Although this mechanism suggests that the fracture process is initiated by a highly localized plastic failure at the microscopic level, the embrittlement manifests macroscopically by lowering ductility as during a tensile test. These researchers have also reported occurrence of HELP in a wide variety of materials having different crystal structures and have asserted that this mechanism is able to explain in general the effect of hydrogen on deformation (both elastic and plastic) and fracture. Several investigators have reported the occurrence of hydrogen embrittlement in hydrogen environment at a pressure of 1 atm or less [6,10]. In these conditions, the pressure of hydrogen within the micro-voids or greater than the ambient cannot be sustained on thermodynamic considerations. In general, the decrease in UTS of steels because of hydrogen is far greater than 1 atm. When the hydrogen charged tensile specimens are tested in air, any signicant amount of pressure induced by the presence of hydrogen in the matrix cannot be sustained within the micro-pores present in the matrix. Such results do not support the gaseous pressure induced model. Precipitation of hydrogen at internal defects to explain IHE is inconsistent with the present results as well since we observe an increase in UTS on hydrogenation. As regards the Petch Stables model of surface adsorption [50], the hydrogen present in the lattice can, in principle, migrate to the pre-existing microvoid/pores, and lower the surface energy. Such microvoids, acting as sites for crack initiation can indeed lower the rupture strength. The fractographic features of both hydrogenated and hydrogen-free materials of this study are essentially ductile and involve considerable plastic deformation before fracture and thus the process of hydrogen adsorption is not responsible for hydrogen embrittlement. Next, we consider the de-cohesion

model in which the decrease in rupture strength is brought about by the weakening of lattice bonds. The absence of cleavage fracture in hydrogen charged indicates that hydrogen is not affecting the lattice energy. The dislocation transport theory [52] suggests that dislocations are capable of generating a large concentration of hydrogen in excess of lattice solubility in some microstructural features. The excess hydrogen may precipitate as high pressure gaseous hydrogen in microvoids and aid void growth which leads to embrittlement. Louthan et al. [53] also showed that localized hydrogen concentration as high as 104 times the equilibrium value may be realised in micro cracks produced by dislocation interactions. This high concentration of hydrogen may create internal gas pressure of the order of 106 atm or more which is large enough to cause embrittlement by pressure expansion mechanism. If the micro-voids/cracks get activated in the presence of hydrogen as postulated in both these mechanisms, then a reduction in the stress for initiation and growth of cracks is expected. In such a situation, UTS should decrease with increasing hydrogen content. In contrast to this, we note that UTS increases with increasing charging time (i.e. increasing hydrogen content). Thus, it appears that both these models do explain the results of this study. For similar reasons, our observations on the effect of hydrogen on strength and work hardening also preclude Beachems [54] and Lynchs [55] models as possible mechanisms of embrittlement. The HELP mechanism of hydrogen embrittlement postulates that hydrogen in solution in the lattice enhances mobility of dislocations by decreasing barriers to dislocation motion and the fracture is a highly localized plastic failure process rather than an embrittlement [56,57]. It must be mentioned here that this is the only mechanism which attempts to rationalize the effects of hydrogen on elastic, plastic and fracture properties on the basis of the interaction of hydrogen with dislocations. It has been shown that H-induced slip localization may also lead to an increase in ow stress depending on the conditions of hydrogen charging and tests variables employed during deformation. The observed effects of hydrogen on the strength, ductility and fracture process of the steels studied appear to be in accordance with the proposition of HELP mechanism. However, the increase in the tangent modulus and the work hardening rate by hydrogen as observed in this investigation cannot be explained if hydrogen decreases the interaction energy of dislocation with other defects. A decrease in interaction energy of dislocations due to solute hydrogen would decrease the elastic modulus [58] and intuitively decrease the rate of work hardening also. Thus the suggestion that hydrogen facilitates motion of dislocation as described in the HELP model is incongruous and against the increase in tangent modulus and work hardening reported in this

280

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281

work. In addition, a decrease rather than an increase in dimple size of both the steels on hydrogenation further substantiates our view that lattice softening as well as HELP are not supported by our results on the effects of hydrogen on deformation behaviour of the two commercial steels. It is difcult to arrive at rm conclusions regarding the exact role of hydrogen in the embrittlement of these steels on the basis of the present results. However, the combination of the effect of hydrogen on mechanical properties and fractographic information does imply that the intrinsic effect of hydrogen is to enhance strength and decrease ductility by impeding the motion of dislocations. From a review of a large number of studies, Louthan et al. [28] have inferred that although there is no agreement among the investigators over the mechanisms of hydrogen embrittlement, all mechanisms suggest accumulation of a highly localized hydrogen concentration as a prerequisite for the embrittlement process. A similar observation has also been made by Pressouyre [32] and he proposed the trap theory which in some respects attempted to unify various other theories. According to this theory, regardless of the mechanism of embrittlement, a crack is initiated or its growth is assisted when the concentration of hydrogen trapped in a pre-existing stressed defect exceeds some critical value. This theory can be used for rationalizing the observations of this study. The important defects in this context are dislocations, particle matrix interfaces, the grain boundaries and magnetic domain walls [59,60]. The association of hydrogen with these structural features inuences deformation behaviour and fracture as discussed earlier. The results of this investigation clearly suggest that some critical stress concentration besides a critical hydrogen concentration must be achieved at some trapping sites in order to nucleate a void for the ductile fracture. In this study, such stress concentration is achieved primarily at the particle matrix interface by dislocation pile up. As all the defects present in the microstructure trap a considerable amount of hydrogen before commencement of the tensile test, further enrichment by large-scale dislocation transport appears to be of only secondary importance in the present context. However, this discussion is somewhat speculative at this stage and further work is needed to conrm this. Thus the observed effects of hydrogen on these two steels may be construed as demonstrating that the increase in strength, work hardening and modulus and loss of ductility were due to the presence of a large concentration of trapped hydrogen in various types of defects introduced by the novel method of charging. Furthermore, an interval of 24 h between charging and tensile testing allows the hydrogen to distribute itself at various trapping sites within the matrix. Hence large-scale redistribution of hydrogen through lattice diffusion may not be taking place during deformation and fracture.

5. Conclusions A novel technique for charging hydrogen has been demonstrated. This method permits hydrogenation to a very high level without causing damage to the microstructure and facilitates studying the intrinsic effect of hydrogen on the mechanical properties of steel. The intrinsic effect of hydrogen on both the steels studied are the enhancement of yield strength, ultimate tensile strength, tangent modulus, work hardening rate and reduction in ductility. The fracture mode was ductile dimple and hydrogen was found to reduce dimple size compared to hydrogen-free materials. The changes in mechanical properties were primarily due to the presence of a large concentration of hydrogen at various traps like dislocations, grain boundaries and matrix/carbide or precipitate interfaces and the interaction of hydrogen with these traps. The results of this investigation could be rationalized by invoking trap theory and it appears that a critical concentration of hydrogen should be available at some stressed defect to initiate a micro-void/crack. In this investigation particle/matrix interfaces have been identied as possible locations of crack initiation.

References
[1] [2] [3] [4] [5] [6] [7] A.R. Troiano, Trans. Am. Soc. Metals 52 (1960) 54. R.A. Oriani, Annu. Rev. Mater. Sci. 8 (1978) 327. J.P. Hirth, Metal. Trans. 11A (1980) 861. H.H. Gray, ASTM STP 543 (1974) 3. M.R. Louthan Jr, in: I.M. Bernstein, A.W. Thompson (Eds.), Hydrogen in Metals, ASM, Metals Park, OH, 1974, p. 53. H.H. Johnson, in: I.M. Bernstein, A.W. Thompson (Eds.), Hydrogen in Metals, ASM, Metals Park, OH, 1974, p. 35. D.A. Westphal, F.J. Worzala, in: I.M. Bernstein, A.W. Thompson (Eds.), Hydrogen in Metals, ASM, Metals Park, OH, 1974, p. 79. D.O. Nothwood, U. Kosasih, Int. Metal. Rev. 28 (1985) 3323. D.L. Davidson, D. Eylon, Metal. Trans. 11A (1980) 837. I.M. Bernstein, R. Garber, G.M. Pressouyre, in: A.W. Thompson, I.M. Bernstein (Eds.), Effect of Hydrogen on Behaviour of Materials, TMS-AIME, New York, 1976, p. 37. J.K. Lin, R.A. Oriani, Acta Metall. 31 (1983) 1077. Y. Tobe, W.R. Tyson, Scripta Metall. 11 (1977) 849. C. St. John, W.W. Gerberich, Metal. Trans. 4A (1973) 589. I.M. Bernstein, Mater. Sci. Eng. 6 (1970) 1. T.P. Groeneveld, A.R. Elsea, ASTM STP 543 (1974) 11. M.A.V. Devanathan, Z.O.J. Stacheurski, Proc. R. Soc. A270 (1962) 90. J.J. Deluccia, D.A. Berman, ASTM STP 227 (1981) 256. J. Volkl, G. Alefeld, in: G. Alefeld, J. Volkl (Eds.), Hydrogen in Metals I (Basic Properties), Springer, Berlin, 1978, p. 321. M.L. Hill, E.W. Johnson, Trans. AIME 221 (1961) 622. J.L. Waisman, G. Sines, L.B. Robinson, Metal. Trans. 4 (1973) 291. C.E. Ells, C.J. Simpson, in: I.M. Bernstein, A.W. Thompson Jr (Eds.), Hydrogen in Metals, ASM, Metals Park, OH, 1974, p. 345.

[8] [9] [10]

[11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21]

G.P. Tiwari et al. / Materials Science and Engineering A286 (2000) 269281 [22] G. Ohma, G.P. Tiwari, Y. Injima, K. Hirano, in: Hydrogen in Metals (JIMIS 2), Minakami, Japan, 1979, p. 229. [23] H.A. Wriedt, R.A. Oriani, Acta Metall. 18 (1970) 753. [24] J.C.M. Li, Y.T. Chou, Trans. AIME 245 (1969) 607. [25] J.M. Hyzak, D.E. Rawl Jr, M.R. Luthan Jr, Scripta Metall. 15 (1981) 937. [26] A.W. Thompson, Mater. Sci. Eng. 14 (1974) 253. [27] A.J. West Jr, M.R. Louthan Jr, Metall. Trans. 13 (1982) 2049. [28] M.R. Louthan, M.J. Morgan, J. Nondestructive Eval. 15 (1996) 113. [29] R.A. Oriani, P.H. Josephic, Metal. Trans. 11A (1980) 1809. [30] H. Kimura, H. Matsui, T. Kimura, in: Hydrogen in Metals (JIMIS 2), Minakami, Japan, 1979, p. 533. [31] P. Bastein, P. Azou, Compd. Rond. 231 (1959) 147. [32] G.M. Pressouyre, in: C.G. Interrante, G.M. Pressuouyre (Eds.), Current Solutions to Hydrogen Problems in Steels, ASM, Metals Park, OH, 1982, p. 18. [33] H.H. Podgurski, R.A. Oriani, Metal. Trans. 3 (1972) 2055. [34] A.J. Kumnick, H.H. Johnson, Acta Metall. 18 (1980) 33. [35] P. Schiller, Nuovo Cimento 276 (1976) 3313. [36] R. Gibala, Trans. AIME 239 (1967) 1574. [37] T.S. Ke, in: Hydrogen in Metals (JIMIS2), Minakami, Japan, 1979, p. 573. [38] H. Kornmuller, in: G. Alefeld, J. Volkel (Eds.), Hydrogen in Metals, Topics Appl. Phys., 28 (1978) 289. [39] C. Wert, O. Buck, in: Hydrogen in Metals (JIMIS 2), Minakami, Japan, 1979, p. 221. [40] J.P. Laurent, G. Lapasset, M. Aucouturier, P. Lacombe, in: Hydrogen in Metals (JIMIS2), Minakami, Japan, 1979, p. 559. [41] R.A. Oriani, P.A. Josephic, Metal. Trans. 11A (1980) 1809.

281

[42] R.D. McCright, in: Proc. Conf. Stress Corrosion Cracking and Hydrogen Embrittlement of Iron Base Alloys, NACE-5 UnieuxFirminy, France, 1973, p. 306. [43] G.M. Pressouyre, I.M. Bernstein, Acta Metall. 27 (1979) 89. [44] J. Smith, in: N.R. Moody, A.W. Thompson (Eds.), Hydrogen Effects on Materials Behaviour, Met. Soc. AIME, Warrendale, PA, 1990. [45] H.K. Birnbaum, Proc. Conf. Environment-Induced Cracking of Metals, NACE, Houston, TX, 1990, p. 21. [46] C.A. Zapffe, A.G. Moore, Trans. AIME 154 (1943) 355. [47] A.R. Troiano, Trans. ASM 52 (1960) 54. [48] R.A. Oriani, Ber. Bunsenges Phys. Chem. 76 (1972) 848. [49] X. Chen, T. Foecke, M. Lii, Y. Katz, W.W. Gerberich, Eng. Fract. Mech. 35 (1990) 997. [50] N.J. Petch, P. Stables, Nature 169 (1952) 842. [51] P. Bastien, P. Azou, in: Proceedings of the 1st World Metallurgical Congress, ASM, Cleveland, OH, 1951. [52] J.K. Tien, A.W. Thompson, I.M. Bernstein, R.J. Richards, Metal. Trans. 7A (1976) 821. [53] M.R. Louthan, G.R. Caskey Jr, J.A. Donovan, D.E. Rawl Jr, Mater. Sci. Eng. 10 (1972) 357. [54] C.D. Beachem, Metall. Trans. 3 (1972) 437. [55] S.P. Lynch, Acta Metall. 36 (1988) 2639. [56] H.K. Birnbaum, P. Sofronis, Mater. Sci. Eng. A176 (1994) 191. [57] P. Sofronis, H.K. Birnbaum, J. Mech. Phys. Solids 43 (1995) 49. [58] P.J. Ferreira, I.M. Robertson, H.K. Birnbaum, Acta Mater. 46 (1998) 1749. [59] R.A. Oriani, Acta Metall. 18 (1970) 147. [60] M. Kurkela, G.S. Frankel, R.M. Latanision, S. Suresh, R.O. Ritchie, Scripta Metal. 16 (1982) 455.

You might also like