Hydrogen embrittlement

Last updated
Hydrogen-induced cracking (HIC) Steel-with-Hydrogen-Induced-Cracks-01.jpg
Hydrogen-induced cracking (HIC)

Hydrogen embrittlement (HE), also known as hydrogen-assisted cracking or hydrogen-induced cracking (HIC), is a reduction in the ductility of a metal due to absorbed hydrogen. Hydrogen atoms are small and can permeate solid metals. Once absorbed, hydrogen lowers the stress required for cracks in the metal to initiate and propagate, resulting in embrittlement. Hydrogen embrittlement occurs most notably in steels, as well as in iron, nickel, titanium, cobalt, and their alloys. Copper, aluminium, and stainless steels are less susceptible to hydrogen embrittlement. [1] [2] [3] [4]

Contents

The essential facts about the nature of hydrogen embrittlement have been known since the 19th century. [5] [6] Hydrogen embrittlement is maximised at around room temperature in steels, and most metals are relatively immune to hydrogen embrittlement at temperatures above 150 °C. [7] Hydrogen embrittlement requires the presence of both atomic ("diffusible") hydrogen and a mechanical stress to induce crack growth, although that stress may be applied or residual. [2] [8] [9] Hydrogen embrittlement increases at lower strain rates. [1] [2] [10] In general, higher-strength steels are more susceptible to hydrogen embrittlement than mid-strength steels. [11]

Metals can be exposed to hydrogen from two types of sources: gaseous hydrogen and hydrogen chemically generated at the metal surface. Gaseous hydrogen is molecular hydrogen and does not cause embrittlement though it can cause hot hydrogen attack (see below). It is the atomic hydrogen from chemical attack which causes embrittlement because the atomic hydrogen dissolves quickly into the metal at room temperature. [6] Gaseous hydrogen is found in pressure vessels and pipelines. Electrochemical sources of hydrogen include acids (as may be encountered during pickling, etching, or cleaning), corrosion (typically due to aqueous corrosion or cathodic protection), and electroplating. [1] [2] Hydrogen can be introduced into the metal during manufacturing by the presence of moisture during welding or while the metal is molten. The most common causes of failure in practice are poorly-controlled electroplating or damp welding rods.

Hydrogen embrittlement as a term can be used to refer specifically to the embrittlement that occurs in steels and similar metals at relatively low hydrogen concentrations, or it can be used to encompass all embrittling effects that hydrogen has on metals. These broader embrittling effects include hydride formation, which occurs in titanium and vanadium but not in steels, and hydrogen-induced blistering, which only occurs at high hydrogen concentrations and does not require the presence of stress. [10] However, hydrogen embrittlement is almost always distinguished from high temperature hydrogen attack (HTHA), which occurs in steels at temperatures above 400 °C and involves the formation of methane pockets. [12] The mechanisms (there are many) by which hydrogen causes embrittlement in steels are not comprehensively understood and continue to be explored and studied. [1] [13] [14]

Mechanisms

Crack in a hardened steel due to hydrogen, observed by scanning electron microscopy (SEM). Intergranular Crack SEM Micrograph.jpg
Crack in a hardened steel due to hydrogen, observed by scanning electron microscopy (SEM).

Hydrogen embrittlement is a complex process involving a number of distinct contributing micro-mechanisms, not all of which need to be present. The mechanisms include the formation of brittle hydrides, the creation of voids that can lead to high-pressure bubbles, enhanced decohesion at internal surfaces and localised plasticity at crack tips that assist in the propagation of cracks. [14] There is a great variety of mechanisms that have been proposed [14] and investigated as to the cause of brittleness once diffusible hydrogen has been dissolved into the metal. [6] In recent years, it has become widely accepted that HE is a complex, material and environmental dependent process, so that no single mechanism applies exclusively. [15]

Material susceptibility

Hydrogen embrittles a variety of metals including steel, [19] [20] aluminium (at high temperatures only [21] ), and titanium. [22] Austempered iron is also susceptible, though austempered steel (and possibly other austempered metals) displays increased resistance to hydrogen embrittlement. [23] NASA has reviewed which metals are susceptible to embrittlement and which only prone to hot hydrogen attack: nickel alloys, austenitic stainless steels, aluminium and alloys, copper (including alloys, e.g. beryllium copper). [2] Sandia has also produced a comprehensive guide. [24]

Steels

Steels were embrittled with hydrogen through cathodic charging. Heat treatment (baking) was used to reduce hydrogen content. Lower bake times resulted in quicker fracture times due to higher hydrogen content. Hydrogen Belittlement.png
Steels were embrittled with hydrogen through cathodic charging. Heat treatment (baking) was used to reduce hydrogen content. Lower bake times resulted in quicker fracture times due to higher hydrogen content.

Steel with an ultimate tensile strength of less than 1000 MPa (~145,000 psi) or hardness of less than HRC 32 on the Hardness Rockwell Scale is not generally considered susceptible to hydrogen embrittlement. As an example of severe hydrogen embrittlement, the elongation at failure of 17-4PH precipitation hardened stainless steel was measured to drop from 17% to only 1.7% when smooth specimens were exposed to high-pressure hydrogen [2]

As the strength of steels increases, the fracture toughness decreases, so the likelihood that hydrogen embrittlement will lead to fracture increases. In high-strength steels, anything above a hardness of HRC 32 may be susceptible to early hydrogen cracking after plating processes that introduce hydrogen. They may also experience long-term failures anytime from weeks to decades after being placed in service due to accumulation of hydrogen over time from cathodic protection and other sources. Numerous failures have been reported in the hardness range from HRC 32-36 and more above; therefore, parts in this range should be checked during quality control to ensure they are not susceptible.

Testing the fracture toughness of hydrogen-charged, embrittled specimens is complicated by the need to keep charged specimens very cold, in liquid nitrogen, to prevent the hydrogen diffusing away. [26]

Copper

Copper alloys which contain oxygen can be embrittled if exposed to hot hydrogen. The hydrogen diffuses through the copper and reacts with inclusions of Cu
2
O
, forming 2 metallic Cu atoms and H2O (water), which then forms pressurized bubbles at the grain boundaries. This process can cause the grains to literally be forced away from each other, and is known as steam embrittlement (because steam is directly produced inside the copper crystal lattice, not because exposure of copper to external steam causes the problem).

Vanadium, nickel, and titanium

Alloys of vanadium, nickel, and titanium have a high hydrogen solubility, and can therefore absorb significant amounts of hydrogen. This can lead to hydride formation, resulting in irregular volume expansion and reduced ductility (because metallic hydrides are fragile ceramic materials). This is a particular issue when looking for non-palladium-based alloys for use in hydrogen separation membranes. [18]

Fatigue

While most failures in practice have been through fast failure, there is experimental evidence that hydrogen also affects the fatigue properties of steels. This is entirely expected given the nature of the embrittlement mechanisms proposed for fast fracture. [27] [16] In general hydrogen embrittlement has a strong effect on high-stress, low-cycle fatigue and very little effect on high-cycle fatigue. [2] [24]

Environmental embrittlement

Hydrogen embrittlement is a volume effect: it affects the volume of the material. Environmental embrittlement [2] is a surface effect where molecules from the atmosphere surrounding the material under test are adsorbed onto the fresh crack surface. This is most clearly seen from fatigue measurements where the measured crack growth rates [24] can be an order of magnitude higher in hydrogen than in air. That this effect is due to adsorption, which saturates when the crack surface is completely covered, is understood from the weak dependence of the effect on the pressure of hydrogen. [24]

Environmental embrittlement is also observed to reduce fracture toughness in fast fracture tests, but the severity is much reduced compared with the same effect in fatigue [24]

Hydrogen embrittlement is the effect where a previously embrittled material has low fracture toughness whatever atmosphere it is tested in. Environmental embrittlement is the effect when the low fracture toughness is only observed when the testing happens in that atmosphere.

Sources of hydrogen

During manufacture, hydrogen can be dissolved into the component by processes such as phosphating, pickling, electroplating, casting, carbonizing, surface cleaning, electrochemical machining, welding, hot roll forming, and heat treatments.

During service use, hydrogen can be dissolved into the metal from wet corrosion or through misapplication of protection measures such as cathodic protection. [2] In one case of failure during construction of the San Francisco–Oakland Bay Bridge galvanized (i.e. zinc-plated) rods were left wet for 5 years before being tensioned. The reaction of the zinc with water introduced hydrogen into the steel. [28] [29] [30]

A common case of embrittlement during manufacture is poor arc welding practice, in which hydrogen is released from moisture, such as in the coating of welding electrodes or from damp welding rods. [22] [31] To avoid atomic hydrogen formation in the high temperature plasma of the arc, welding rods have to be perfectly dried in an oven at the appropriate temperature and duration before use. Another way to minimize the formation of hydrogen is to use special low-hydrogen electrodes for welding high-strength steels.

Apart from arc welding, the most common problems are from chemical or electrochemical processes which, by reduction of hydrogen ions or water, generate hydrogen atoms at the surface, which rapidly dissolve in the metal. One of these chemical reactions involves hydrogen sulfide ( H
2
S
) in sulfide stress cracking (SSC), a significant problem for the oil and gas industries. [32]

After a manufacturing process or treatment which may cause hydrogen ingress, the component should be baked to remove or immobilize the hydrogen. [29]

Prevention

Hydrogen embrittlement can be prevented through several methods, all of which are centered on minimizing contact between the metal and hydrogen, particularly during fabrication and the electrolysis of water. Embrittling procedures such as acid pickling should be avoided, as should increased contact with elements such as sulfur and phosphate.

If the metal has not yet started to crack, hydrogen embrittlement can be reversed by removing the hydrogen source and causing the hydrogen within the metal to diffuse out through heat treatment. This de-embrittlement process, known as low hydrogen annealing or "baking", is used to overcome the weaknesses of methods such as electroplating which introduce hydrogen to the metal, but is not always entirely effective because a sufficient time and temperature must be reached. [33] Tests such as ASTM F1624 can be used to rapidly identify the minimum baking time (by testing using design of experiments, a relatively low number of samples can be used to pinpoint this value). Then the same test can be used as a quality control check to evaluate if baking was sufficient on a per-batch basis.

In the case of welding, often pre-heating and post-heating the metal is applied to allow the hydrogen to diffuse out before it can cause any damage. This is specifically done with high-strength steels and low alloy steels such as the chromium/molybdenum/vanadium alloys. Due to the time needed to re-combine hydrogen atoms into the hydrogen molecules, hydrogen cracking due to welding can occur over 24 hours after the welding operation is completed.

Another way of preventing this problem is through materials selection. This will build an inherent resistance to this process and reduce the need of post processing or constant monitoring for failure. Certain metals or alloys are highly susceptible to this issue, so choosing a material that is minimally affected while retaining the desired properties would also provide an optimal solution. Much research has been done to catalog the compatibility of certain metals with hydrogen. [24] Tests such as ASTM F1624 can also be used to rank alloys and coatings during materials selection to ensure (for instance) that the threshold of cracking is below the threshold for hydrogen-assisted stress corrosion cracking. Similar tests can also be used during quality control to more effectively qualify materials being produced in a rapid and comparable manner.

Surface Coatings

Coatings act as a barrier between the metal substrate and the surrounding environment, hindering the ingress of hydrogen atoms. These coatings can be applied through various techniques such as electroplating, chemical conversion coatings, or organic coatings. The choice of coating depends on factors such as the type of metal, the operating environment, and the specific requirements of the application.

Electroplating is a commonly used method to deposit a protective layer onto the metal surface. This process involves immersing the metal substrate into an electrolyte solution containing metal ions. By applying an electric current, the metal ions are reduced and form a metallic coating on the substrate. Electroplating can provide an excellent protective layer that enhances corrosion resistance and reduces the susceptibility to hydrogen embrittlement.

Chemical conversion coatings are another effective method for surface protection. These coatings are typically formed through chemical reactions between the metal substrate and a chemical solution. The conversion coating chemically reacts with the metal surface, resulting in a thin, tightly adhering protective layer. Examples of conversion coatings include chromate, phosphate, and oxide coatings. These coatings not only provide a barrier against hydrogen diffusion but also enhance the corrosion resistance of the metal.

Organic coatings, such as paints or polymer coatings, offer additional protection against hydrogen embrittlement. These coatings form a physical barrier between the metal surface and the environment. They provide excellent adhesion, flexibility, and resistance to environmental factors. Organic coatings can be applied through various methods, including spray coating, dip coating, or powder coating. They can be formulated with additives to further enhance their resistance to hydrogen ingress.

Thermally sprayed coatings offer several advantages in the context of hydrogen embrittlement prevention. The coating materials used in this process are often composed of materials with excellent resistance to hydrogen diffusion, such as ceramics or cermet alloys. These materials have a low permeability to hydrogen, creating a robust barrier against hydrogen ingress into the metal substrate. [34]

Testing

Most analytical methods for hydrogen embrittlement involve evaluating the effects of (1) internal hydrogen from production and/or (2) external sources of hydrogen such as cathodic protection. For steels, it is important to test specimens in the lab that are at least as hard (or harder) than the final parts will be. Ideally, specimens should be made of the final material or the nearest possible representative, as fabrication can have a profound impact on resistance to hydrogen-assisted cracking.

There are numerous ASTM standards for testing for hydrogen embrittlement:

There are many other related standards for hydrogen embrittlement:

Notable failures from hydrogen embrittlement

See also

Related Research Articles

<span class="mw-page-title-main">Electroplating</span> Creation of protective or decorative metallic coating on other metal with electric current

Electroplating, also known as electrochemical deposition or electrodeposition, is a process for producing a metal coating on a solid substrate through the reduction of cations of that metal by means of a direct electric current. The part to be coated acts as the cathode of an electrolytic cell; the electrolyte is a solution of a salt of the metal to be coated; and the anode is usually either a block of that metal, or of some inert conductive material. The current is provided by an external power supply.

<span class="mw-page-title-main">Corrosion</span> Gradual destruction of materials by chemical reaction with its environment

Corrosion is a natural process that converts a refined metal into a more chemically stable oxide. It is the gradual deterioration of materials by chemical or electrochemical reaction with their environment. Corrosion engineering is the field dedicated to controlling and preventing corrosion.

<span class="mw-page-title-main">Brazing</span> Metal-joining technique

Brazing is a metal-joining process in which two or more metal items are joined by melting and flowing a filler metal into the joint, with the filler metal having a lower melting point than the adjoining metal.

Plating is a finishing process in which a metal is deposited on a surface. Plating has been done for hundreds of years; it is also critical for modern technology. Plating is used to decorate objects, for corrosion inhibition, to improve solderability, to harden, to improve wearability, to reduce friction, to improve paint adhesion, to alter conductivity, to improve IR reflectivity, for radiation shielding, and for other purposes. Jewelry typically uses plating to give a silver or gold finish.

Sulfide stress cracking (SSC) is a form of hydrogen embrittlement which is a cathodic cracking mechanism. It should not be confused with the term stress corrosion cracking which is an anodic cracking mechanism. Susceptible alloys, especially steels, react with hydrogen sulfide (H2S), forming metal sulfides (MeS) and atomic hydrogen (H) as corrosion byproducts. Atomic hydrogen either combines to form H2 at the metal surface or diffuses into the metal matrix. Since sulfur is a hydrogen recombination poison, the amount of atomic hydrogen which recombines to form H2 on the surface is greatly reduced, thereby increasing the amount of diffusion of atomic hydrogen into the metal matrix. This aspect is what makes wet H2S environments so severe.

<span class="mw-page-title-main">Stress corrosion cracking</span> Growth of cracks in a corrosive environment

Stress corrosion cracking (SCC) is the growth of crack formation in a corrosive environment. It can lead to unexpected and sudden failure of normally ductile metal alloys subjected to a tensile stress, especially at elevated temperature. SCC is highly chemically specific in that certain alloys are likely to undergo SCC only when exposed to a small number of chemical environments. The chemical environment that causes SCC for a given alloy is often one which is only mildly corrosive to the metal. Hence, metal parts with severe SCC can appear bright and shiny, while being filled with microscopic cracks. This factor makes it common for SCC to go undetected prior to failure. SCC often progresses rapidly, and is more common among alloys than pure metals. The specific environment is of crucial importance, and only very small concentrations of certain highly active chemicals are needed to produce catastrophic cracking, often leading to devastating and unexpected failure.

<span class="mw-page-title-main">Intergranular fracture</span>

Intergranular fracture, intergranular cracking or intergranular embrittlement occurs when a crack propagates along the grain boundaries of a material, usually when these grain boundaries are weakened. The more commonly seen transgranular fracture, occurs when the crack grows through the material grains. As an analogy, in a wall of bricks, intergranular fracture would correspond to a fracture that takes place in the mortar that keeps the bricks together.

<span class="mw-page-title-main">Intergranular corrosion</span> When crystallite boundaries are more corrosive than their interiors

In materials science, intergranular corrosion (IGC), also known as intergranular attack (IGA), is a form of corrosion where the boundaries of crystallites of the material are more susceptible to corrosion than their insides.

Hydrogen damage is the generic name given to a large number of metal degradation processes due to interaction with hydrogen atoms. Note that molecular gaseous hydrogen does not have the same effect as atoms or ions released into solid solution in the metal.

In materials science, environmental stress fracture or environment assisted fracture is the generic name given to premature failure under the influence of tensile stresses and harmful environments of materials such as metals and alloys, composites, plastics and ceramics.

Liquid metal embrittlement is a phenomenon of practical importance, where certain ductile metals experience drastic loss in tensile ductility or undergo brittle fracture when exposed to specific liquid metals. Generally, a tensile stress, either externally applied or internally present, is needed to induce embrittlement. Exceptions to this rule have been observed, as in the case of aluminium in the presence of liquid gallium. This phenomenon has been studied since the beginning of the 20th century. Many of its phenomenological characteristics are known and several mechanisms have been proposed to explain it. The practical significance of liquid metal embrittlement is revealed by the observation that several steels experience ductility losses and cracking during hot-dip galvanizing or during subsequent fabrication. Cracking can occur catastrophically and very high crack growth rates have been measured.

<span class="mw-page-title-main">Chromate conversion coating</span> Chemical treatment of metals

Chromate conversion coating or alodine coating is a type of conversion coating used to passivate steel, aluminium, zinc, cadmium, copper, silver, titanium, magnesium, and tin alloys. The coating serves as a corrosion inhibitor, as a primer to improve the adherence of paints and adhesives, as a decorative finish, or to preserve electrical conductivity. It also provides some resistance to abrasion and light chemical attack on soft metals.

<span class="mw-page-title-main">Embrittlement</span> Loss of ductility of a material, making it brittle

Embrittlement is a significant decrease of ductility of a material, which makes the material brittle. Embrittlement is used to describe any phenomena where the environment compromises a stressed material's mechanical performance, such as temperature or environmental composition. This is oftentimes undesirable as brittle fracture occurs quicker and can much more easily propagate than ductile fracture, leading to complete failure of the equipment. Various materials have different mechanisms of embrittlement, therefore it can manifest in a variety of ways, from slow crack growth to a reduction of tensile ductility and toughness.

The salt spray test is a standardized and popular corrosion test method, used to check corrosion resistance of materials and surface coatings. Usually, the materials to be tested are metallic and finished with a surface coating which is intended to provide a degree of corrosion protection to the underlying metal.

Electrogalvanizing is a process in which a layer of zinc is bonded to steel in order to protect against corrosion. The process involves electroplating, running a current of electricity through a saline/zinc solution with a zinc anode and steel conductor. Such Zinc electroplating or Zinc alloy electroplating maintains a dominant position among other electroplating process options, based upon electroplated tonnage per annum. According to the International Zinc Association, more than 5 million tons are used yearly for both hot dip galvanizing and electroplating. The plating of zinc was developed at the beginning of the 20th century. At that time, the electrolyte was cyanide based. A significant innovation occurred in the 1960s, with the introduction of the first acid chloride based electrolyte. The 1980s saw a return to alkaline electrolytes, only this time, without the use of cyanide. The most commonly used electrogalvanized cold rolled steel is SECC, acronym of "Steel, Electrogalvanized, Cold-rolled, Commercial quality". Compared to hot dip galvanizing, electroplated zinc offers these significant advantages:

Low hydrogen annealing, commonly known as "baking" is a heat treatment in metallurgy for the reduction or elimination of hydrogen in a material to prevent hydrogen embrittlement. Hydrogen embrittlement is the hydrogen-induced cracking of metals, particularly steel which results in degraded mechanical properties such as plasticity, ductility and fracture toughness at low temperature. Low hydrogen annealing is called a de-embrittlement process. Low hydrogen annealing is an effective method compared to alternatives such as electroplating the material with zinc to provide a barrier for hydrogen ingress which results in coating defects.

Mechanical plating, also known as peen plating, mechanical deposition, or impact plating, is a plating process that imparts the coating by cold welding fine metal particles to a workpiece. Mechanical galvanization is the same process, but applies to coatings that are thicker than 0.001 in (0.025 mm). It is commonly used to overcome hydrogen embrittlement problems. Commonly plated workpieces include nails, screws, nuts, washers, stampings, springs, clips, and sintered iron components.

Metallurgical failure analysis is the process to determine the mechanism that has caused a metal component to fail. It can identify the cause of failure, providing insight into the root cause and potential solutions to prevent similar failures in the future, as well as culpability, which is important in legal cases. Resolving the source of metallurgical failures can be of financial interest to companies. The annual cost of corrosion in the United States was estimated by NACE International in 2012 to be $450 billion a year, a 67% increase compared to estimates for 2001. These failures can be analyzed to determine their root cause, which if corrected, would save reduce the cost of failures to companies.

Rising Step Load Testing is a testing system that can apply loads in tension or bending to evaluate hydrogen-induced cracking. It was specifically designed to conduct the accelerated ASTM F1624 step-modified, slow strain rate tests on a variety of test coupons or structural components. It can also function to conduct conventional ASTM E8 tensile tests; ASTM F519 200-hr Sustained Load Tests with subsequent programmable step loads to rupture for increased reliability; and ASTM G129 Slow Strain Rate Tensile tests.

Metal-induced embrittlement (MIE) is the embrittlement caused by diffusion of metal, either solid or liquid, into the base material. Metal induced embrittlement occurs when metals are in contact with low-melting point metals while under tensile stress. The embrittler can be either solid (SMIE) or liquid. Under sufficient tensile stress, MIE failure occurs instantaneously at temperatures just above melting point. For temperatures below the melting temperature of the embrittler, solid-state diffusion is the main transport mechanism. This occurs in the following ways:

References

  1. 1 2 3 4 Lynch, S. P. (2011-01-01), Raja, V. S.; Shoji, Tetsuo (eds.), "2 - Hydrogen embrittlement (HE) phenomena and mechanisms", Stress Corrosion Cracking, Woodhead Publishing Series in Metals and Surface Engineering, Woodhead Publishing, pp. 90–130, ISBN   978-1-84569-673-3 , retrieved 2022-06-10
  2. 1 2 3 4 5 6 7 8 9 NASA (2016). "Hydrogen Embrittlement" (PDF). Retrieved 18 December 2020.{{cite journal}}: Cite journal requires |journal= (help)
  3. Jewett, R. P.; Walter, R. J.; Chandler, W. T.; Frohmberg, R. P. (1973-03-01). "Hydrogen environment embrittlement of metals".{{cite journal}}: Cite journal requires |journal= (help)
  4. Safety Standard for Hydrogen and Hydrogen Systems: Guidelines for Hydrogen System Design, Materials Selection, Operations, Storage, and Transportation (PDF). Vol. NSS 1740.16. Washington, DC: Office of Safety and Mission Assurance, National Aeronautics and Space Administration. 1997-10-29. p. A-93. Retrieved 2022-06-27.
  5. Johnson, William H. (31 December 1875). "II. On some remarkable changes produced in iron and steel by the action of hydrogen and acids" (PDF). Proceedings of the Royal Society of London. 23 (156–163): 168–179. doi:10.1098/rspl.1874.0024. eISSN   2053-9126. ISSN   0370-1662. JSTOR   113285. S2CID   97579399.
  6. 1 2 3 Bhadhesia, Harry. "Prevention of Hydrogen Embrittlement in Steels" (PDF). Phase Transformations & Complex Properties Research Group, Cambridge University. Retrieved 17 December 2020.
  7. "What is hydrogen embrittlement? – Causes, effects and prevention". TWI - The Welding Institute. Retrieved 18 December 2020.
  8. Oriani, R A (August 1978). "Hydrogen Embrittlement of Steels". Annual Review of Materials Science. 8 (1): 327–357. Bibcode:1978AnRMS...8..327O. doi:10.1146/annurev.ms.08.080178.001551. ISSN   0084-6600.
  9. "Hydrogen Embrittlement". Metallurgy for Dummies. Retrieved 18 December 2020.
  10. 1 2 Louthan, M. R. (2008-06-01). "Hydrogen Embrittlement of Metals: A Primer for the Failure Analyst". Journal of Failure Analysis and Prevention. 8 (3): 289–307. doi:10.1007/s11668-008-9133-x. ISSN   1864-1245. S2CID   51738408.
  11. Li, Hanyu; Niu, Ranming; Li, Wei; Lu, Hongzhou; Cairney, Julie; Chen, Yi-Sheng (September 2022). "Hydrogen in pipeline steels: Recent advances in characterization and embrittlement mitigation". Journal of Natural Gas Science and Engineering. 105: 104709. doi:10.1016/j.jngse.2022.104709. S2CID   250713252.
  12. TWI – The Welding Institute. "What is high temperature hydrogen attack (HTHA) / hot hydrogen attack?". TWI - The Welding Institute. Retrieved 16 December 2020.
  13. 1 2 3 Barnoush, Afrooz. "Hydrogen embrittlement revisited by in situ electrochemical nanoindentations" (PDF). Archived from the original (PDF) on 2011-05-18. Retrieved 18 December 2020.
  14. 1 2 3 Robertson, Ian M.; Sofronis, P.; Nagao, A.; Martin, M. L.; Wang, S.; Gross, D. W.; Nygren, K. E. (2015). "Hydrogen Embrittlement Understood". Metallurgical and Materials Transactions A. 46A (6): 2323–2341. Bibcode:2015MMTA...46.2323R. doi: 10.1007/s11661-015-2836-1 . S2CID   136682331.
  15. 1 2 Haiyang Yu (February 2009). "Discrete dislocation plasticity HELPs understand hydrogen effects in bcc materials". Journal of the Mechanics and Physics of Solids. 123: 41–60. arXiv: 1807.05101 . doi: 10.1016/j.jmps.2018.08.020 . S2CID   56081700.
  16. 1 2 Vergani, Laura; Colombo, Chiara; et al. (2014). "Hydrogen effect on fatigue behavior of a quenched and tempered steel". Procedia Engineering. 74 (XVII International Colloquium on Mechanical Fatigue of Metals (ICMFM17)): 468–71. doi: 10.1016/j.proeng.2014.06.299 .
  17. Song, Jun (11 November 2012). "Atomic mechanism and prediction of hydrogen embrittlement in iro". Nature Materials. 12 (2): 145–151. doi:10.1038/nmat3479. PMID   23142843 . Retrieved 22 December 2020.
  18. 1 2 Dolan, Michael D.; Kochanek, Mark A.; Munnings, Christopher N.; McLennan, Keith G.; Viano, David M. (February 2015). "Hydride phase equilibria in V–Ti–Ni alloy membranes". Journal of Alloys and Compounds. 622: 276–281. doi:10.1016/j.jallcom.2014.10.081.
  19. Djukic, M.B.; et al. (2014). "Hydrogen embrittlement of low carbon structural steel". Procedia Materials Science. 3 (20th European Conference on Fracture): 1167–1172. doi: 10.1016/j.mspro.2014.06.190 .
  20. Djukic, M.B.; et al. (2015). "Hydrogen damage of steels: A case study and hydrogen embrittlement model". Engineering Failure Analysis. 58 (Recent case studies in Engineering Failure Analysis): 485–498. doi:10.1016/j.engfailanal.2015.05.017.
  21. Ambat, Rajan; Dwarakadasa (February 1996). "Effect of Hydrogen in aluminium and aluminium alloys: A review". Bulletin of Materials Science. 19 (1): 103–114. doi: 10.1007/BF02744792 .
  22. 1 2 Eberhart, Mark (2003). Why Things Break. New York: Harmony Books. p.  65. ISBN   978-1-4000-4760-4.
  23. Tartaglia, John; Lazzari, Kristen; et al. (March 2008). "A Comparison of Mechanical Properties and Hydrogen Embrittlement Resistance of Austempered vs Quenched and Tempered 4340 Steel". Metallurgical and Materials Transactions A. 39 (3): 559–76. Bibcode:2008MMTA...39..559T. doi:10.1007/s11661-007-9451-8. ISSN   1073-5623. S2CID   136866718.
  24. 1 2 3 4 5 6 Marchi, C. San (2012). "Technical Reference for Hydrogen Compatibility of Materials" (PDF).
  25. Morlet, J. G. (1958). "A new concept in hydrogen embrittlement in steels". The Journal of the Iron and Steel Institute. 189: 37.
  26. Fracture Mechanics Techniques for Assessing the Effects of Hydrogen on Steel PropertiesM J Cheaitani and R J Pargeter, TWI, paper presented at the International Steel and Hydrogen Conference 28 September 2011.
  27. Fernandez-Sousa, Rebeca (2020). "Analysis of the influence of microstructural traps on hydrogen assisted fatigue". Acta Materialia. 199: 253. arXiv: 2008.05452 . Bibcode:2020AcMat.199..253F. doi:10.1016/j.actamat.2020.08.030. S2CID   221103811.
  28. 1 2 Francis, Rob. "A Failure Analysis of Hydrogen Embrittlement in Bridge Fasteners". Corrosionpedia. Retrieved 18 December 2020.
  29. 1 2 Ferraz, M. Teresa; Oliveira, Manuela (2008). "Steel fasteners failure by hydrogen embrittlement" (PDF). Ciência e Tecnologia dos Materiais. 20 (1/2): 128–133. Retrieved 18 December 2020.
  30. 1 2 Yun Chung (2 December 2014). "Validity of Caltrans' Environmental Hydrogen Embrittlement Test on Grade BD Anchor Rods in the SAS Span" (PDF).
  31. Weman, Klas (2011). Welding Processes Handbook. Elsevier. p. 115. ISBN   978-0-85709-518-3.
  32. "Standard Test Method for Process Control Verification to Prevent Hydrogen Embrittlement in Plated or Coated Fasteners". Astm.org. Retrieved 24 February 2015.
  33. Federal Engineering and Design Support. "Embrittlement" (PDF). Fastenal. Fastenal Company Engineering Department. Retrieved 9 May 2015.
  34. "ADDRESSING HYDROGEN PERMEATION AND EMBRITTLEMENT". 2023.
  35. "ASTM F1459 - 06(2012): Standard Test Method for Determination of the Susceptibility of Metallic Materials to Hydrogen Gas Embrittlement (HGE)". Astm.org. Retrieved 2015-02-24.
  36. "ASTM G142 - 98(2011) Standard Test Method for Determination of Susceptibility of Metals to Embrittlement in Hydrogen Containing Environments at High Pressure, High Temperature, or Both". Astm.org. Retrieved 2015-02-24.
  37. ASTM STP 543, "Hydrogen Embrittlement Testing"
  38. Raymond L (1974). Hydrogen Embrittlement Testing. ASTM International. ISBN   978-0-8031-0373-3.
  39. "ASTM F1940 - 07a(2014) Standard Test Method for Process Control Verification to Prevent Hydrogen Embrittlement in Plated or Coated Fasteners". Astm.org. Retrieved 2015-02-24.
  40. "ASTM F519 - 17a Standard Test Method for Mechanical Hydrogen Embrittlement Evaluation of Plating/Coating Processes and Service Environments". www.astm.org. Retrieved 21 April 2018.
  41. Mair, Lucy (14 January 2015). "British Land to replace 'a number of bolts' on Leadenhall Building". constructionnews.co.uk. Retrieved 21 April 2018.
  42. "Cheesegrater bolts to cost Severfield £6m after Leadenhall building loses five". cityam. 17 June 2015. Retrieved 22 December 2020.