See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/270230762 Effects of Shielding with Various Hydrogen-Argon Mixtures on Supermartensitic Stainless Steel TIG Welds Article in Materials Testing · May 2010 DOI: 10.3139/120.110135 CITATIONS READS 5 217 3 authors, including: Dan Eliezer Thomas Kannengiesser Ben-Gurion University of the Negev Bundesanstalt für Materialforschung und -prüfung 590 PUBLICATIONS 6,863 CITATIONS 181 PUBLICATIONS 1,541 CITATIONS SEE PROFILE SEE PROFILE Some of the authors of this publication are also working on these related projects: HELIUM IN METALS View project [Materials] Special Issue "Technologies for Joining and Forming Thin-Walled Structures in the Construction of Transportation Vehicles" IF=3.057 View project All content following this page was uploaded by Dan Eliezer on 06 June 2017. The user has requested enhancement of the downloaded file. MP MATERIALS TESTING IN JOINING TECHNOLOGY Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. Effects of Shielding with Various Hydrogen-Argon Mixtures on Supermartensitic Stainless Steel TIG Welds 306 Dan Eliezer, Y. Nissim, Beer-Sheva, Israel, and Thomas Kannengießer, Berlin A number of common defects in stainless steel welding result from the presence of hydrogen in the weld. The service life of the stainless steel joints is further significantly dependent on the presence of hydrogen in the respective environment and the susceptibility of various weld microstructures to hydrogen degradation. As a relatively new materials generation, supermartensitic stainless steels (SMSS) are increasingly applied to substitute more expensive alloys, particularly in the oil and gas industries. As a result of their martensitic microstructure these alloys are prone to hydrogen assisted cracking (HAC). The resistance of SMSS to hydrogen assisted stress corrosion cracking (HASCC) during sour service has been extensively studied, predominantly for industrial purposes. Studies are primarily conducted with parent materials based on standard test procedures. The principal hydrogen behavior in welded SMSS microstructures has been less investigated. The central objectives of the study are to determine the hydrogen interactions with the microstructure of a Gas Tungsten Arc (GTA) welded SMSS and hydrogen trapping mechanisms. The interactions of hydrogen with various Tungsten Inert Gas (TIG) welded SMSS microstructures are investigated by means of X-ray diffraction (XRD) and optic (OM) and electronic microscopy (SEM). A number of methods have been employed for the estimation of the quantities of absorbed hydrogen. Hydrogen interaction with structural defects and the characteristics of hydrogen desorption have been studied by means of thermal desorption spectroscopy (TDS), and hydrogen content measurements (LECO analyses). The effects of the respective microstructure on hydrogen absorption and desorption behavior are discussed in detail. Supermartensitic stainless steels are essentially iron-based alloys which contain chromium, nickel, and molybdenum. They contain approximately the same amount of chromium as conventional martensitic stainless steels of the AISI 410 and 420 series, but their carbon content is reduced. The addition of up to 6.5 wt.-% nickel substantially decreases the martensite transformation towards lower temperatures; weld microstructures may contain considerable amounts of retained and annealing austenite. Three different levels of nickel and molybdenum, which were commercially available, were therefore selected for fitness purposes. Lean alloys are intended to be used in sweet service and contain nickel below 2 wt.-% and molybdenum below 1 wt.-%, and are fabricated as low © Carl Hanser Verlag, München MP Materials Testing 52 (2010) 5 Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MATERIALS TESTING IN JOINING TECHNOLOGY carbon (LC) steels with carbon contents below 0.02 wt.-% and with nickel and molybdenum amounts of between 2 and 4 wt.-% and of approximately 1 wt.-%, respectively, medium alloys have been developed for intermediate sour service applications. At increasing H2S levels and decreasing pH extra low carbon (ELC) fat grades are applied. These materials contain even less carbon, i. e. under 0.01 wt.-%, but even more nickel, up to 6.5 wt.-%, and more molybdenum up to 2.5 wt.-% [1-4]. The base materials are usually heat treated in several stages and consist of considerable portions of annealing and retained austenite, depending on the nickel content [2]. Supermartensitic alloys have been commercially available since the early 1990’s and have been increasingly applied, almost exclusively in the oil & gas industries. In these industries the selection of materials which fit exactly with the intended applications has become increasingly important for economic reasons. For example, supermartensitic stainless steels represent an attractive cost reducing alternative to higher alloys and more expensive materials, and are thus increasingly applied as flowline materials at several North Sea oil and gas fields. At these locations, welded components are subjected to sour service conditions, providing a potential risk for hydrogen uptake, degradation and cracking of supermartensitic stainless steels. As another potential failure risk for welded SMSS, Hydrogen Assisted Cold Cracking (HACC) has to be avoided during welding. Even small hydrogen concentrations, trapped inside the material, might lead to failure. Consequently, a comprehensive knowledge of hydrogentrapping interactions is necessary to make any decision and/or judgment as to whether a trap site or a particular trapped hydrogen content is detrimental to the safe service operation of welded joints of such materials. Several industrial application-oriented studies have been performed to investigate the sour service and respective HASCC resistance of SMSS, predominantly in the as-delivered, but also in the welded condition. Limited investigations have further been undertaken to evaluate their cold cracking resistance. A preliminary comparison revealed the same fracture topography SMSS weld metals for HACC and for HASCC [1]. 52 (2010) 5 C S P Mn Si Ni Cr Mo N 0.06 0.0009 0.021 1.870 0.294 6.498 11.65 2.330 0.009 MP Table 1. Chemical composition of the SMSS (wt.-%) Therefore, the objective of the present study is to investigate hydrogen interactions with GTA welded supermartensitic stainless steel microstructures in greater detail. Particularly with respect to adsorption and desorption characteristics, as well as to hydrogen trapping mechanisms. Experimental Approach The chemical composition of the tested supermartensitic stainless steel is listed in Table 1. In order to investigate hydrogen affects on the microstructure, as well as on the absorption and desorption behavior of hydrogen in supermartensitic stainless steel, the alloy was exposed to hydrogen in three different modes. These are as follows: electrochemical hydrogenation, gaseous-phase hydrogenation and hydrogen introduced by GTA welding via a mixed Ar + H2 shielding gas. Hydrogen charging was carried out electrochemically (cathodic charging) at room temperature. The cathodic charging technique generates H+ on the sample’s surface. A power supply (galvanostat) impresses a constant current on the electrolytic cell, such that the specimen is cathodic relative to an inert electrode such as platinum (Pt). The anode, a platinum wire, is located symmetrically to the cathodic polarized specimen, thereby uniformly distributing the potential around the specimen. The specimens were cathodically hydrogenated at identical charging times, using a current density of 50 mA/cm2. The charging was carried out in 1N D2SO4 + 0.25 g NaAsO2 electrolyte. Gaseous phase charging was carried out in H2 UHP 99.99 atmospheres, using a pressure of 1 atm during 6.5 hours at room temperature and 650 °C. The charging of the weld metal with hydrogen was performed by a Gas Tungsten Arc Welding (GTAW) using the parameters summarized in Table 2. Prior to the electrochemical and gas phase hydrogen exposure, samples were cut from SMSS plates, and were mechanically polished up to 0.05 μ. The total content of hydrogen/deuterium absorbed in the alloy was measured by a LECO RH-404 hydrogen determination system, using the test protocol described in the previous literature [5]. Hydrogen/Deuterium evolution and trapping characteristics were studied by thermal desorption spectroscopy (TDS), a technique that involves measurement of the desorption rate of gas atoms, soluted or trapped in the material, while heating the sample at a known rate [6-7]. After ultrasonic cleaning in ethanol, the deuterium charged specimens were placed in the specimen holder; the system was sealed and pumped down to 10 mPa. The heating rates (ramps) were 3, 5, and 7 K min-1, and the temperature range was between 25-600 °C, both parameters being programmed into a temperature controller. While heating, the mass spectrometer was placed in a continuous mode for scanning atomic masses in the range of between 3.5-4.5 amu. Shielding Gas Type Shielding Gas [l/min] Current [A] Voltage [V] Weld Speed [cm/min] Argon + 0% hydrogen 10 145 10 30 Argon + 2% hydrogen 10 145 11.5 32 Argon + 5% hydrogen 10 145 11.5 32 Argon + 7.5% hydrogen 10 145 11.5 39 Table 2. Welding parameters of SMSS 307 Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MP MATERIALS TESTING IN JOINING TECHNOLOGY Figure 1. Hydrogen contents in different welding areas of SMSS TIG welded in gas shield atmospheres containing different hydrogen percentages Specimens for optical microscopy were prepared by mechanical polishing through 0.05 μm alumina. The samples were finally etched using one of the following etchings: 1. 5 gr. cupric chloride in 100 cc hydrochloric acid + 100 cc ethyl alcohol + 100 cc water 2. 10 gr. ferric chloride in 20 cc hydrochloric acid + 4 cc nitric acid. The use of different etchings was necessary to highlight specific features and phases and thus obtain complementary information on the microstructures. Digital photographs were taken using a Leica DMR digital camera. The specimen’s surface was observed using a JEOL JSM 5600 scanning electron micro- scope (SEM). The machine was operated at an accelerating voltage of 15 kV. The microstructure and phase characterization of the above specimens were studied by means of X-Ray diffraction (XRD) analyses, using a Rigaku Type 2000 X-ray powder diffractometer with a CuKα radiation (λ = 1.54059 Å) and a graphite monochromator diffracted beam. The data was collected in an angular range of 41° < 2θ < 54°, 61° < 2θ < 94° by steps of 0.05° (2θ) with a constant counting time of 4s by step. Results and Discussion Hydrogen Absorption. The amount of hydrogen absorbed in the alloy in the Sample As-received 4.06 ± 0.28 Charged for 24 h with deuterium 165 ± 11.55 Gas phase hydrogenation (R.T., 6.5 hours, 1 atm) 2.16 ± 0.11 Gas phase hydrogenation (650 °C, 6.5 hours, 1 atm) 1.87 ± 0.15 TIG welded in Ar 0 % H2 6.7 ± 0.34 TIG welded in Ar 2 % H2 6.2 ± 0.42 TIG welded in Ar 5 % H2 5.83 ± 0.25 TIG welded in Ar 7.5 % H2 8.64 ± 0.60 Table 3. Hydrogen content in SMSS alloy 308 Absorbed Hydrogen (ppm) as-received state and after exposure to hydrogen was measured by means of a LECO analysis. Results are summarized in Table 3. The hydrogen contents in different welding areas of the TIG welded specimens welded with gas shield atmospheres containing different hydrogen percentages are presented in Figure 1. It needs to be noted that in its asdelivered condition the base material contains about 4 ppm. Table 3 demonstrates the different ways by which the hydrogen was introduced to the alloy significantly influence the amount of absorbed hydrogen. It can be generally stated that the higher the fugacity of hydrogen in the atmosphere the larger the hydrogen concentrations absorbed in the alloy. The hydrogen content absorbed in the electrochemically charged specimen is the highest and differs at a rate of about two orders of magnitude from that of the as-received specimen. This result can be attributed to the high hydrogen concentrations produced on the sample surface, suppressing most of the recombination during the electrochemical charging period, as reported in former studies [2]. The distribution of hydrogen in the volume is dependent upon the hydrogen charging technique, i. e. if hydrogen is introduced from the surface, as occurs during electrolytic or gaseous-phase charging, or is introduced by a welding arc and homogenously distributed, as occurs in a weld metal. The internal hydrogen distribution is also significantly dependent on the various weld microstructures, due to their different diffusivities and solubility rates [2, 8]. It should be noted that the hydrogen amount absorbed during gas-phase hydrogenation is considerably low, and even lower than in the as-delivered base material. This is to be attributed to prior degassing in the process. From the above mentioned result, which shows some consistency with a similar experimental study with carbon mild steels [9-10], it can tentatively be concluded that at low pressures hydrogen will not enter the material from a gaseous phase even at elevated temperatures of up to 650 °C. This is as long as the material is not considerably plastically deformed under a static or dynamic mechanical load. From Figure 1, it can be seen that the different weld microstructures significantly influence the amount of hydrogen 52 (2010) 5 Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MATERIALS TESTING IN JOINING TECHNOLOGY absorbed in the specimen, due to their variable solubility rates. The lowest concentrations have been measured in the heat affected zones (HAZ) which are to be attributed to the heat introduced into the HAZ during welding. This result stands in contrast to a recent investigation of the HAZ multilayer GTA welded SMSS of similar chemical composition where the highest or at least the similarly high hydrogen concentrations have been found in the HAZ as compared to the heat affected zone. As shown in Figure 2, such high hydrogen levels in the HAZ probably have to be attributed to a high fraction of retained or annealing austenite in the HAZ of about 35 %. Such annealing and austenite forming effects are not present in the mono-layer weld investigated here, and therefore less hydrogen is accumulated in the HAZ. Also, in contrast to the previous study with multilayer welds [2], somewhat higher hydrogen concentrations than in the HAZ have been found in the weld metal of the mono-layer weld investigated. The hydrogen content in the weld metal exceeds that of the base material only if very high hydrogen percentages are added in the shielding gas, i. e. at a fraction of 7.5 % in argon. Hydrogen’s Effects on the Microstructure. The exposure parameters extensively influence the amount, quality and depth of hydrogen absorption. This means that as a consequence of the hydrogenation process different microstructural morphologies might be obtained. Also, the respective phase transformations are dependent on the amount and way in which hydrogen is introduced into the material. For instance, hydrogen represents an interstitial, and might shift the martensite formation towards lower temperatures during the cooling period after welding. The initial microstructure and XRD pattern of the studied material are shown in Figures 3 and 4. The phase analysis of SMSS XRD patterns reveals the existence of two phases of BCCmartensite and FCC-austenite arranged in a lamellar structure. The γ-phase reflections exhibits significant broadening, but the FCC lattice parameter remains unchanged. The α-phase reflections, on the other hand, exhibits negligible broadening and the BCC lattice parameter contracts by approximately 0.8 %. This result can be attributed to the fact that hydrogen 52 (2010) 5 MP Figure 2. Solubility evaluated by the sub-surface concentration of hydrogen dependent on the austenite content in the weld microstructures of similar nickel and molybdenum alloyed SMSS multi-layer steel welds [2] Figure 3. X-Ray diffraction patterns of electrochemically charged SMSS alloy with deuterium for 24 hours Figure 4. Typical microstructure of the as-received SMSS, a) optical micrograph, b) SEM micrograph 309 Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MP MATERIALS TESTING IN JOINING TECHNOLOGY Figure 5. X-Ray diffraction patterns of TIG welds in gas shield atmospheres containing Ar +7.5 % hydrogen, from different welding areas Figure 6. Optical micrographs of the SMSS TIG weld microstructures shielded with Ar +7.5 % hydrogen, a) weld metal and HAZ at the fusion line, b) HAZ, c) weld metal, d) weld metal has very different diffusion rates in austenitic and in martensitic microstructures. The rate is higher in body centered cubic lattices, and remains high even at low temperatures. Therefore, the retained and continuous γ phase in the fully lamellar microstructure of the SMSS alloy offers a blocking site for hydrogen diffusion. The microstructure and XRD pattern of the different weld microstructures are shown in the Figures 5 and 6 above. The 310 phase analysis of SMSS XRD patterns reveals the existence of two phases BCCmartensite and FCC-austenite according to the WRC 1992 diagram [11]. As to be expected from such a mono-layer weld, the base material contains a larger amount of retained or annealing austenite in comparison to the HAZ area, while the weld metal completely transformed into martensite. Figure 7 exhibits X-Ray diffraction patterns of TIG welded SMSS alloys, from the weld metal, at different hydrogen percentages in the atmosphere. As the hydrogen percentage in the atmosphere increases, more hydrogen is introduced into the completely transformed martensitic lattice and consequently, the α-phase reflections exhibits significant broadening. Results of XRD patterns after gas-phase hydrogenation are shown in Figure 8. Evidently the soluted hydrogen concentrations under these charging conditions are too small to cause any broadening of the γ- and α-phase reflections, i. e. no lattice parameter expansion occurs at such conditions. Desorption Charcteristics. In contrast to conventional hydrogen measurement and extraction technologies used for welding applications, such as the mercury procedure or carrier gas hot extraction, the thermal desorption spectroscopy (TDS) technique represents a much more suitable, reliable, and expressive method to characterize the hydrogen evolution process, and to assess hydrogen trapping characteristics in a microstructure. For this reason it has been utilized within this study. The spectra analyses were supported by data obtained from a variety of other experimental techniques, such as LECO hydrogen quantity analyses, XRD, and microstructure investigations by means of optic and electronic microscopy [5-6]. The trapping phenomena in steels are not well understood and their connection with hydrogen assisted cracking is unclear. For instance, it is unclear whether hydrogen deeply trapped within a Titanium Carbide may not be (re-)activated during the straining of the microstructure thereby contributing to cracking. Furthermore, although it has been shown and it can be anticipated that hydrogen solubility increases and its diffusivity decreases with the increase of chromium carbide precipitation in SMSS [2] it is not evident that these hydrogen accumulations contribute to inter-granular cracking [1]. At the same rate, it is however difficult to obtain unambiguous experimental information on the influence of traps for cracking, partly as a result of the small hydrogen concentration involved. Yet there is a generally consensus in past literature that the lattice diffusivity of hydrogen in steel is strongly retarded by the presence of microstructural inhomogeneities including grain boundaries, 52 (2010) 5 Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MATERIALS TESTING IN JOINING TECHNOLOGY dislocations, carbides, and nonmetallic particles, all acting as potential hydrogen trapping sites [12-14]. The capacity of a microstructural feature to act as a trapping site is governed by the potential-energy well (and binding energy (Eb)) relative to the normal interstitial positions in the Fe-lattice. With the increasing strength of the trapping site relative to the energy well depth, i.e. the peak height of the free energy graph, traps are categorized as reversible and then irreversible. In the absence of traps, hydrogen diffusion occurs by a random walk between interstitial lattice positions, usually with a very short residential time for the diffusing of a hydrogen atom. Hydrogen diffusion can be described by the ideal lattice diffusion coefficient, DL, and governed by the activation energy for diffusion associated with the energy barrier (on lattice migration energy (Em)) between such interstitial positions E ––m–– [13], i. e. DL = D0 · e––R·T . The activation energy in this study is calculated by the Lee and Lee model [7], which is based on the following assumptions: (a) the controlling process is a first order, which is described by detrapping and diffusion to the surface, i. e. the chemical absorption energy (Echem) is smaller than the migration energy (Em) or smaller than the binding energy (Eb), and second order desorption phenomena, such as molecular desorption accompanied by recombination of adsorbed atoms, are disregarded (b) with the increase of the heating rate of the specimens the peaks are shifted to higher temperatures. Taking into account the shift of the peaks at different heating rates, the effective energy of hydrogen desorption was evaluated from the slope. Characteristics of hydrogen desorption from the SMSS specimens are presented in Figures 9 to 11. The calculated results from these TDS graphs are summarized in Tables 4 to 6. For all of the SMSS specimens subjected to hydrogen in the “as- received”, TIG welded, and gas phase hydrogenated condition, the TDS spectra exhibit a common feature that initial degassing occurs until the vacuum reaches 10-7 to 10-8 mbar. Furthermore, all the spectra exhibited a first peak at relatively low temperatures, i. e. around room temperature, indicating a larger loss of hydrogen. These peaks are probably associated with hy- 52 (2010) 5 MP Figure 7. X-Ray diffraction patterns in the base material of the TIG welded SMSS at different hydrogen percentages in the atmosphere drogen escaping from the specimen’s surface at a relatively low activation energy of approximately 4.5 kJ. A second evolution peak occurs in all spectra over a wide temperature range from about 80 °C up to 200 °C with an activation energy of approximately 5.7-6.7 kJ. The spectra for the as-received material (Figure 9a), for the 1 atm gaseouscharged material (Figure 9b), for the weld metal, and the base material produced without hydrogen in the shielding gas (Figure 10a) and finally for the base material of the weld shielded with 7,5 vol.-% hydrogen in the argon atmosphere (Figure 10b) – all have nearly identical shapes. As to be expected, no additional hydrogen was taken up in the material during welding without hydrogen. Also, no hydrogen was traced in the base material during welding with 7.5 vol.-% in the shielding gas. Most importantly, almost no additional hydrogen was taken up during charging at room-temperature in a gaseous hydrogen environment. However, concerning the 7.5 vol.-% hydrogen shielded weld metal and after gas-phase hydrogenation at high temperature, it was observed that hydrogen evolution reaches a characteristic peak at approximately 285-300 °C, and covers a wide temperature range between Figure 8. X-Ray diffraction patterns (1 atm, 6.5 hours) of SMSS after gas-phase hydrogenation, charged at room temperature and 650 °C 311 MATERIALS TESTING IN JOINING TECHNOLOGY Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MP a) b) Figure 9. TDS spectrum of SMSS at different charging temperatures and at a heating rate of 3 °C/min a) as-received material, b) gas-phase charged (1 atm, 6.5 hours) a) b) Figure 10. TDS spectrum for different SMSS welded microstructures of alloys, TIG welded in different shielding gas atmospheres at a heating rate of 3 °C/min, a) Ar +0 % H2 , b) Ar +7.5 % H2 a) b) Figure 11. TDS spectrum of samples electrochemically charged with deuterium for 24 hours, TIG welded in gas shield atmospheres containing Ar +7.5 % H2 at a heating rate of 3 °C/min, a) as-received material, b) weld metal ΔT = 140 °C and 360 °C with an activation energy of approximately 8.5 to 9.1 kJ. In comparison to the literature, the activation energy values are calculated close to migration energy: pure annealed Fe, Em is 7 kJ/mol (0.07 eV/atom), and DL is 1.3 · 10-5 cm2/s at 25 °C [2]. 312 The specimens subjected to electrochemical hydrogen charging exhibited completely different hydrogen evolution characteristics. In particular, a large peak was observed at approximately 150 °C that covers a wide range of temperature between ΔT = 25 °C and 400 °C. It needs to be emphasized that such a peak occurs after hydrogenation of both the “as-delivered” material and the weld metal previously welded in a 7.5 vol.-% hydrogen-argon atmosphere (Figure 11 a and b). 52 (2010) 5 Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. MATERIALS TESTING IN JOINING TECHNOLOGY A comparison of the TDS plots of the electrochemically hydrogenated specimens (Figure 10) with those of other specimens (Figures 8 and 9) reveals that the amount of deuterium trapped inside the material and desorbed at this temperature is higher than that of other specimens. Preliminary metallographic investigations reveal the existence of different inhomogenities in the martensitic microstructure of these welds, and it is to be expected that these defects act as various trapping sites. For example, the largest activation energy value obtained from the respective desorption peak calculations stands at approximately 9 kJ/mole, indicating that the respective trap belongs to a H-dislocation elastic stress field. With respect to the various potential trapping sites in supermartensitic stainless steel weld microstructures, the following considerations should be taken: a) Desorption energy is defined as the sum between the trap binding energy and the activation energy for lattice diffusion (Ed = Eb + Em). b) When the activation energy for lattice diffusion (Em) is very large relative to the activation or binding energy of a trap (Eb), a diffusion controlled hydrogen evolution will be measured. As discussed above, the theoretical and measured values of the activation energy are numerically very close. This in turn means that bulk diffusion is not negligible and therefore that Lee and Lee’s theory [7] does not apply in this case. Lee and Lee’s theoretical approach, which is based on hydrogen evolution from trap sites only, is attractive because of its theoretical simplicity. However Lee and Lee’s model [7] does not account for diffusion, correspondingly it further does not account for the possibility of hydrogen released from one trap becoming available to other traps. Lee and Lee’s model is therefore unsuitable for the explanation of hydrogen evolution and for the calculation of the activation energy for the investigated supermartensitic stainless steel weld microstructures. Conclusions The investigation of the affects of different hydrogen exposures to welded supermartensitic stainless steels has lead to a better understanding of microstructural influences on the absorption/ 52 (2010) 5 Specimen Peak # Temp. at desorption peak (°C) Maximal desorption rate (x1014 Hatoms/gr · s) Half height peak width (°C) Calculated activation energy (kJ) I 30 11.30 50.78 4.2 II 92.34 3.39 37.50 5.1 III 130 2.80 25.78 6.5 7.90 51.23 4.56 MP As-Received As-Received Ar + 0 % H2 Base Metal Weld Metal I 32.3 II 88.11 3.19 30.20 5.72 III 121 2.81 31.50 6.65 IV 143.85 2.63 28.43 6.68 I 30.94 8.74 50.31 4.54 II 82.34 3.40 28.59 5.70 III 110.94 2.94 27.97 6.62 IV 162.19 2.57 202.50 6.70 7.91 42.34 4.56 Ar + 7.5 % H2 Base Metal Weld Metal I 32.5 II 89.38 3.19 30.00 5.72 III 120 2.72 37.50 6.66 IV 143.75 2.57 28.59 6.68 I 34.84 9.33 47.03 4.57 II 92.5 3.56 36.56 5.73 III 117.5 3.24 27.19 6.63 IV 172.34 3.42 108.91 6.70 V 302.34 3.65 137.19 9.10 Table 4. Parameters of thermal desorption analysis from TIG welded SMSS, at a heating rate of 3 °C/min Peak # Temp. at desorption peak (°C) Maximal desorption rate (·1014 D-atoms/g · s) Half height peak width (°C) As-Received charged with deuterium for 24 hours I 77.5 112 38.44 II 155 167 172.50 Weld metal of TIG welded in Ar+7.5 % H2 charged with deuterium for 24 hours I 80 128 40.63 II 152.34 195 148.13 Table 5. Parameters of thermal desorption analysis from SMSS electrochemically charged with deuterium, for 24 hours at a heating rate of 3 °C/min desorption behavior and the interaction between hydrogen atoms and possible trap sites. The following conclusions can therefore be drawn: 1. The amount of absorbed hydrogen is significantly dependent on the different modes of hydrogenation. Electrochemical charging can lead to much higher levels of absorbed hydrogen concentrations than charging in a gaseous hydrogen atmosphere or during welding SMSS in a hydrogenated shielding gas with up to 7.5 vol.-% H2. 2. The absorbed hydrogen concentration significantly depends on the weld microstructure. In the investigated 313 MP MATERIALS TESTING IN JOINING TECHNOLOGY Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. mono-layer weld, lower concentrations are soluted in the HAZ as compared to the weld metal and to the base material. This is to be attributed to a respectively higher content of retained and annealing austenite generated during production of the base material and also to a higher amount of retained austenite in the weld metal itself. However, as the re-heating and annealing processes produce respectively more austenite in the HAZ, much higher hydrogen concentrations can be taken up in the HAZ multilayer joints. The level of hydrogen in the weld microstructures is thus also dependent on the joint design. 3. The influence of hydrogen rates in the atmosphere on absorbed amounts appear only in atmospheres containing 7.5 % H2. Although there is some hydrogen loss in the period between welding and final extraction procedures, if the atmosphere contains a higher amount of hydrogen, more hydrogen could be absorbed in the material and be trapped there. 4. The changing of the lattice parameter depends on the way in which hydrogen is introduced into the weld microstructures. For example, the lattice parameters of the γ- and α-phase remain unchanged after hydrogenation from a gaseous environment. However, the large amounts of hydrogen introduced during electrochemical charging procedures significantly reduce the α-phase reflections, while the parameter of the γ-phase remains unchanged due to the higher solubility of this phase. 5. The hydrogen evolution process was found to be dependent on the hydrogenation mode. For specimens that were subjected to hydrogen in the atmosphere (as-received, TIG welded, and gas phase hydrogenation) TDS spectrums exhibited several evolution peaks. For specimens that were subjected to electrochemical hydrogenation, hydrogen evolution was observed in a large single peak that covered a wide temperature range but was located at lower temperatures. This means that hydrogen is trapped more deeply in the investigated SMSS weld microstructures if introduced during welding rather than during other introductory procedures such as cathodic charging. 314 Specimen R.T. 650 °C Peak # Temp. at desorption peak (°C) Maximal desorption rate (x1014 Hatoms/gr · s) Half height peak width (°C) Calculated activation energy (kJ) I 40.2 6.84 39.34 4.56 II 92.6 2.13 29.29 5.73 III 120.1 1.98 28.18 6.64 I 40 7.64 39.84 4.55 II 92.5 2.84 29.69 5.73 III 120 2.21 28.28 6.64 IV 285 2.99 358.28 7.59 Table 6. Parameters of thermal desorption analysis from SMSS, charged at 1 atm for 6.5 hours at a heating rate of 3 °C/min Abstract Auswirkungen des Schützens mit verschiedenen WasserstoffArgon-Gemischen auf WIG-Schweißungen supermartensitischer Stähle. Eine Anzahl verschiedener Effekte ergibt sich aus der Anwesenheit von Wasserstoff während des Schweißens hochlegierter Stähle. Die Betriebsdauer von geschweißten Bauteilen ist außerdem stark von der Anwesenheit von Wasserstoff im Umgebungsmedium und der Anfälligkeit der verschiedenen Schweißnahtgefüge für eine Degradation ihrer Eigenschaften durch Wasserstoff abhängig. Als eine relative neue Werkstoffgeneration finden supermartensitische hoch legierte Stähle (Supermartensitic Stainless Steels – SMSS) zunehmend als Ersatz für teuere Legierungen insbesondere in der Öl- und Gasindustrie Verwendung. Als Konsequenz ihres martensitischen Gefüges sind diese Legierungen anfällig für eine wasserstoffunterstützte Rissbildung (Hydrogen Assisted Cracking – HAC). Der Widerstand von supermartensitischen Stählen gegen wasserstoffunterstützte Spannungsrisskorrosion (Hydrogen Assisted Stress Corrosion Cracking – HASCC) unter Sauergasbedingungen wurde vor allem für industrielle Einsatzzwecke extensiv untersucht. Solche Studien vornehmlich an Grundwerkstoffen basieren überwiegend auf Standard-Prüfverfahren. Dem gegenüber würde das grundsätzliche Verhalten von Wasserstoff in den Gefügen geschweißter supermartensitischer Stähle wenig untersucht. Die zentralen Gründe für die diesem Beitrag zugrunde liegende Studie waren daher, die Effekte des Wasserstoffs auf das Gefüge von Wolfram Inert Gas (WIG)Schweißungen supermartensitischer Stähle und die entsprechenden Wasserstoff-Trapping-Mechanismen zu untersuchen. Die Wirkungen des Wasserstoffs auf die verschiedenen WIG-geschweißten Gefüge wurden mittels Röntgendiffraktometrie, Lichtmikroskopie und Rasterelektronenmikroskopie untersucht. Eine Anzahl von Verfahren wurde außerdem angewendet, um den absorbierten Wasserstoff quantitativ zu bestimmen. Die Wechselwirkung zwischen Wasserstoff mit den mikrostrukturellen Defekten und die Charakteristika der Wasserstoffdesorption wurden mittels Ther52 (2010) 5 MATERIALS TESTING IN JOINING TECHNOLOGY Materials Testing downloaded from www.hanser-elibrary.com by BAM Bundesanstalt fürMaterialprüfung Bibliothek on June 6, 2017 For personal use only. mischer Desorptionsspektroskopie (TDS) und Trägergas-Heißextraktionen des Wasserstoffs (LECO Analyse) untersucht. Die Wirkung des Gefüges auf die Absorption und Desorption von Wasserstoff werden im Detail diskutiert. Further investigation is required, which will use alternative approaches and account for diffusion controlled hydrogen transport in order to demonstrate which trapping sites actually occur in the weld microstructures. References 1 Th. Boellinghaus: Hydrogen assisted cracking in supermartensitic stainless steels, in: N. R. Moody, A. W. Thompson, R. Ricker, G. Was, and R. Jones (Eds.) Hydrogen Effects on Material Behavior and Corrosion Deformation Interactions, TMS, Wyoming (2003), pp. 1009-1018 2 D. Seeger; Th. Boellinghaus: Hydrogen permeation in supermartensitic stainless steel weld microstructures, CORROSION 2004, NACE International, Houston (2004), paper 04142 3 E. Folkhard: Metallurgie der Schweißung hochlegierter Stähle, Springer Verlag Wien (1984), pp. 279-280 (in German) 4 Th. Boellinghaus, H. Hoffmeister: Hydrogen permeation in supermartensitic stainless steels dependent on heat treatment and chemical composition, CORROSION 2000, NACE International, Houston (2000), paper 00141 5 E. Tal-Gutelmacher, D. Eliezer, D. Eylon: The effects of low fugacity hydrogen in duplex- and beta-annealed Ti-6Al-4V alloy, Materials Science and Engineering A 381 (2004) pp. 230-236 6 E. Tal-Gutelmacher; N. Eliaz, D. Eliezer, D. Zander, L. Jastrow, U. Köster: Absorption/desorption behavior of hydrogen and deuterium in a Pd-Coated Zr-based amorphous alloy, Materials Science and Engineering A 358 (2003), pp. 219-225 7 S.-M. Lee; J.-Y. Lee: The trapping and transport phenomena of hydrogen in nickel, in Metallurgical Transactions A, 17 (1986), pp. 181-187 52 (2010) 5 View publication stats 8 P. Rozenak; D. Eliezer: Effects of metallurgical variables on hydrogen embrittlement in AISI Type 316, 321 and 347 stainless steels, Mater. Sci. Eng. 61 (1983), No. 1, pp. 31-41 9 M. Kesten; H. Gräfen: Druckwasserstoffangriff auf unlegierte und niedriglegierte Stähle im Temperaturbereich unterhalb 200 °C, D. Kuron (Ed.): Wasserstoff und Korrosion, 2nd Edition, Bonner Studien Reihe, Bonn (2000), pp. 101-126 10 H. Spähn: Druckwasserstoffangriff auf unlegierte und niedriglegierte Stähle im Temperaturbereich oberhalb 200 °C, D. Kuron (Ed.): Wasserstoff und Korrosion, 2nd Edition, Bonner Studien Reihe, Bonn (2000), pp. 127-183 11 D. J. Kotecki; T. A. Siewert: WRC-1992 constitution diagram for stainless steel weld metals: A modification of the WRC-1998 diagram, Welding Journal, 71 (1992), No. 5, pp. 171-178 12 Th. Boellinghaus, H. Hoffmeister, A. Dangeleit: A scatterband for hydrogen diffusion coefficients in micro-alloyed low carbon structural steels, Welding in the World 35 (1995), No. pp. 83-96 13 P. G. Shewmon: Diffusion in Solids, McGraw-Hill, New York (1960) 14 J. Crank: Mathematics of Difusion, Clarendon Press, Oxford (1970) 15 Th. Boellinghaus; H. Hoffmeister; C. Middel: Scatterbands for hydrogen diffusion coefficients in steels having a ferritic or martensitic microstructure and steels having an austenitic microstructure at room temperature, Welding in the World, 37 (1996), No. 1, pp. 16-23 MP The Authors of This Contribution Prof. Eliezer is a full Professor in the Department of Materials Engineering, Ben Gurion University of the Negev (BGU), Israel, and holds the Eric Samson Chair for Advanced Materials and Processing at the BGU. Prof. Eliezer is a fellow of the ASM (American Society of Materials) (in recognition of distinguished contribution in the field of materials science and materials engineering), and is a Mercator Professor at DFG Clausthal. Prof. Eliezer has held the position of “Visiting Professor” and “Senior Associate” at a number of Universities and Research Institutes across Europe, Asia, and USA, including the: Air Force Wright Aeronautical Laboratories, National Research Council, USA; NASA-Ames Research Centre, National Research Council, USA; University of Illinois, Department of Metallurgy and Mining, USA; Federal Institute for Materials Research and Testing (BAM), Germany, and the Research Centre for Hydrogen Industrial Use and Storage (HYDROGENIUS), Japan. Prof. Eliezer has published over 470 papers in Journals and Conference Proceedings, and has edited numerous Scientific Books and Collective Volumes. Yafit Nissim completed her B. Sc and M. Sc at the Department of Materials Engineering at the Ben-Gurion University of the Negev, Beer-Sheva, Israel. Her M.Sc Thesis “Hydrogen Embrittlement of Welded Stainless Steel” was completed in 2006. Yafit Nissim currently works at Intel Israel. Dr.-Ing. Thomas Kannengießer, born in 1971, studied Mechanical Engineering with the background of Materials Engineering and Materials Testing at the University of Magdeburg. From 1997 to 1999, he was a doctoral candidate at the BAM, Federal Institute for Material Research and Testing, in Berlin and obtained his doctoral degree with the subject “Investigations into the Formation of Welding-Specific Stresses and Deformations at Variable Restraint Conditions in the Component Weld Test”. Since 2005 he is the head of the working group “Component Testing” at BAM. You will find the article and additional material by entering the document number MP110135 on our website at www.materialstesting.de 315