The Truth about the Global Helium Shortage

Article

The Column

Incognito dons his deerstalker to investigate the truth behind the headlines.

Incognito dons his deerstalker to investigate the truth behind the headlines.

Anyone using gas chromatography (GC) in Europe or the US will have noticed that helium (the most popular carrier gas for capillary gas chromatography) prices continue to rise, and most providers are limiting the volumes that can be bought. This, as usual, has led to much speculation about a “helium crisis”.

This is the third “helium crisis” since 2010 (1), and speculation is rife over the future of the carrier gas, leading to many articles in the popular press (2–5) on conversion from helium to hydrogen carrier gas, which can be produced in the laboratory through the electrolytic decomposition of water or pressure swing adsorbtion technology.

I decided to do a little more digging into the supply issues we continue to suffer, in the hope that we can make an informed decision whether to call our local hydrogen or nitrogen gas generator manufacturer. I’ll start by saying that my research has brought forth some widely differing information, however, I believe what is written below represents a true picture of the situation that we face.

 

Helium Sources and Extraction

The main source of helium production is via the radioactive decay of uranium and thorium, which gives rise to alpha particles that subsequently gain electrons to form helium atoms. As shown in the reaction scheme, He-4 is the most abundant isotope.

 

However, the half-life of U-238 is around 4.5 billion years, and given that the Earth is estimated to be 4.7 billion years old, the uranium we have on Earth presently has taken since the dawn of time to produce and accumulate. Put another way, only a vanishingly small amount of “new” helium is produced on an annual basis and therefore, helium is rightly accorded a “non‑renewable resource” status.

The helium-4 formed through underground radioactive decay is typically released into the atmosphere (and subsequently quickly lost into space) through permeable rock; however, where the rock is impermeable, the helium is trapped alongside natural gas deposits. The concentration of the helium in natural gas deposits varies widely from 0.01% to 10% v/v, and a field containing as little as 0.3–0.5% v/v helium is considered economically viable for helium recovery.

Natural gas needs to have key impurities (typically nitrogen, water vapour, carbon dioxide, helium, and other non-combustible materials) removed in order to increase its potential heat energy and enable its use in combustible applications. When the natural gas contains more than around 0.4% v/v helium, it is first sprayed with monoethanolamine to remove CO2 and is then passed through a molecular sieve to remove water vapour followed by activated charcoal to remove any heavy hydrocarbons. This process removes all species that may solidify and plug the cryogenic processing equipment used to extract the helium.

Cryogenic high-pressure fractional distillation is then used to separate helium and nitrogen from methane and subsequently cooled in a condenser to separate the crude helium (which contains 50–70% helium, 1–3% residual methane, small quantities of hydrogen and neon, and the remaining balance is nitrogen).

This crude helium is then cooled to -315 ºF where the nitrogen and remaining methane condense and are drained off. Air is then added and warmed in the presence of a catalyst, causing hydrogen to react with oxygen from the air to form water vapour, which can be condensed through cooling and drawn off.

A pressure swing adsorption (PSA) unit is then employed to separate high purity helium (99.99%) from any remaining impurities by selective removal using microporous adsorbents (6).

Where helium concentrations within the natural gas deposits are low (typically below 0.3% by volume), companies use more advanced, multi-bed PSA technology to recover the helium in an economically viable manner (7).

There is an equilibrium concentration of 5.2 ppm in the Earth’s atmosphere, however, at this level the economics of large-scale recovery (via air liquification) of industrial quality helium are challenging; estimates of 10,000 times more expensive than refining from natural gas sources have been postulated (8).

It should be noted that alternative opinions exist on the source of helium production and there is a school of thought that production is mantle-driven, coming from much deeper in the Earth, rather than being formed in rock deposits nearer the surface, and because it’s so light it’s able to make its way up to the surface where it’s stored with natural gas (9).

Helium must be cooled to almost absolute zero (-425 ºF) if liquefication is required for storage, shipment, or use.

As you will know, there is no way to chemically manufacture helium.

 

Uses of Helium

Figure 1 shows the main uses of helium in the United States (it should be noted that the relative percentage usage figures from other information sources vary significantly).

In magnetic resonance imaging (MRI) medical applications, helium is used as a coolant for the superconducting magnets that are used to produce high resolution images of internal organs. Most of us will also be aware that nuclear magnetic resonance (NMR) laboratory spectrometers also use superconducting magnets, which also require helium for cooling.

Helium is used in large airships and of course for party balloons. It is also used alongside fuel within rockets and to pre-cool liquid hydrogen in the fuel tanks of space vehicles. The Saturn V rocket is estimated to have used 370 thousand cubic metres of helium to launch.

Helium is used extensively in the manufacture of silicon wafers used in semiconductor chip manufacture due to its highly inert nature. It is also used for cooling in quantum computing and is reported to increase the storage capacity of computer hard drives by up to 50% when used to fill the drive housing.

Helium is used as a shielding gas in tungsten arc welding for applications where, at high temperatures, the weld joint will be weakened by the presence of nitrogen or oxygen. Even though the use of argon as a shielding gas would be less expensive, materials with higher heat conductivity, such as aluminium or copper, benefit from the increased heat conductivity of helium during the welding process.

Because helium diffuses through solids three times faster than air, it is used as a tracer gas to detect leaks in high-vacuum equipment (such as cryogenic tanks) and high-pressure containers. The tested object is placed in a chamber, which is then evacuated and filled with helium. Any escaping helium is detected by a sensitive helium mass spectrometer (MS), even at very small leak rates. Handheld devices are also available for this process.

Helium is used to reduce the proportion of nitrogen and oxygen to below those of air in breathing gas mixtures used by deep sea divers. A lower proportion of nitrogen is required to reduce nitrogen narcosis and other physiological effects of nitrogen at depth.

The large hadron collider at Conseil Européen pour la Recherche Nucléaire (CERN) in Switzerland uses 1000 superconducting magnets to bend high energy particles around its 16-mile circumference. In order to maintain a high electrical current with zero electrical resistance, the magnets need to be cooled to -452 ºF. To achieve this, 130 metric tons of liquid helium are required, which is equivalent to 768,000 cubic metres of helium gas.

Of course, of primary interest to readers, is that we use helium as the carrier gas of choice for capillary gas chromatography applications.

 

Geographic Locations and Production

Figure 2 shows estimated major helium resources as of 2018.

The total estimate of global reserves of helium in 2018 were 51.9 billion cubic metres including minor deposits in Poland and Australia (not shown in Figure 2).

The US National (sometimes called Federal) Helium Program Reserve (currently administrated by the US Bureau of Land Management) is held at the Cliffside Storage Facility in a natural underground facility (often called the Bush Dome) located in the Texas Panhandle, near the natural gas fields around Amarillo. It was established in 1925 due to the high helium content (1.9% by volume) of the natural gas fields in the surrounding area. The reserve mainly stores crude helium, which is piped from several fields over many hundreds of miles, most notably from the Bushton reserves located 425 miles away in Kansas.

A further large facility in La Barge, Wyoming, is a major producer of helium accounting for 43% of US helium production and 31% of global production between 2000 and 2012 (10). Estimated production of helium in the US in 2018 is 90 million cubic metres (10).

The Qatari production facility is based in Ras Laffan Industrial City 80 miles north of Doha, and is the largest of Qatar’s liquified natural gas production plants. Ras Laffan plants 1 and 2 are reported to be capable of producing (19.8 + 45.9) 65.7 million cubic metres per annum. A further plant in Qatar (Ras Laffan 3) is due to come on stream in 2021 with a production capability of 12 million cubic metres, bringing Qatar’s production capability to around 77 million cubic metres per annum (11).

The Skikda and Arzew facilities in the north of Algeria have a production capacity of around 8 million cubic metres of helium per annum, although this is regulated to an estimated total production capacity of 16 million cubic metres due to safety concerns after accidents at the facility (12). Recent new investment announcements (2018) were made to significantly increase Algeria’s helium production capabilities, although estimates of the new production capabilities were not available (13).

The Orenburg helium production facility in south-east Russia is estimated to have a production capacity of 8.8 million cubic metres (14), although several new production facilities in Amur and Yaraktinsky with an estimated total production capacity of 67 million cubic metres are planned to come on stream between 2021 and 2024 (15).

In 2016, interest was reinvigorated in very large deposits of helium in the Rift Valley region of Tanzania. This is principally due to the very high concentration of helium, up to 18% by volume, which is delivered by geothermal activity and is typically carried by nitrogen rather than being associated with natural gas deposits. In volcanic areas such as the Rift Valley, the trapped helium can be liberated through weaknesses in the Earth’s crust and bubble up through hot springs. The Tanzania deposits are estimated to contain as much as 3.5 billion cubic metres of helium and the latest reports estimate that production in the region may come on stream as early as 2020 (16).

 

Is There Really a Shortage?

In the early- to mid-20th century helium was considered by the US Government to be of strategic military importance for dirigibles and barrage balloons, and latterly for cooling applications. The Mineral Leasing Act of the 1920s reserved all helium production for US Government use, whilst the Helium Acts Amendment of 1960 allowed the US Bureau of Mines to arrange for five private plants to recover helium from natural gas, effectively a government purchase of the helium stockpile. By 1995 however, the National Helium Reserve was $1.4 billion dollars in debt and the Helium Privatization Act of 1996 mandated that the helium reserve should be sold off in order to pay down these debts. The hope was that by 2015 the reserve would have been sold off and that private industry would establish new sources of helium to satisfy demand. Critically, the pricing for the auctions of the reserve were set below market value in order to ensure that the debt was paid off quickly, keeping global pricing relatively low, however, ongoing costs of production and maintenance of the reserve have led to unpredictable price adjustments up to market levels, affecting the helium price globally. The interest of private industry in helium production did match the 1996 expectations of the US Government, and in 2013 the Helium Stewardship Act was signed into law, allowing sales from the National Helium Reserve to continue to September 2021, effectively ending federal involvement with the reserve at that time. This act effectively slowed the sale of federal helium and allowed the Bureau of Land Management to auction the gas at higher prices. The bill prevented the federal government from undercutting private producers, thus encouraging more sources of helium production to go online.

In June 2017, several countries in the Middle East suspended their trading and political relationships with Qatar, due to its alleged support of terrorism. Qatar distributed its helium through a large port in the United Arab Emirates, which entailed it travelling through Saudi Arabia. Both countries were involved in the embargo of goods from Qatar and therefore production from the Ras Laffan facility was suspended for three weeks and the new distribution route-once supplies became available due to plant recommissioning time-via Oman is more expensive.

The shortages of helium in 2017 caused the Russian Government to embargo all helium production for use within Russian territories.

Whist figures vary widely, here is a summary of the global helium situation as far as I can determine:

Consumption (Estimated):

157 million cubic metres per annum (2016) (17)

157 to 188 million cubic metres estimated for 2019 based on various growth scenarios (18)

Production (Estimated):

160 million cubic metres (2018) (10)

Global Resource (Estimated):

51.9 billion cubic metres (10)

So, by these estimates, we have global reserves for the next 324 years of consumption based on recent usage, indicating that a catastrophic shortage of helium due to exhaustion of helium resources is not a possibility.

Of course, this assumes no growth in helium consumption, which is an unsafe assumption given that current estimates through to 2027 are between 1.5% and 3% compound growth. However, short‑term increases in consumption may well be balanced by the growth in production from Russia, Qatar, and Algeria of a least 79 million cubic metres of helium-50% of current global annual requirements-which are forecast to be online by 2024. These estimates do not include production capability of the geothermal sources in Tanzania.

Whilst production estimates approximately match consumption in 2018, as outlined above, geopolitical factors are responsible for temporary reduction in supply volumes and large fluctuations in price. What is important to us as analytical chemists is a guaranteed supply, and, of course, the price, of helium.

Figure 3 shows one estimate of the supply and demand estimates from 2020 to 2027 (18), and assumes that the supplementary supplies from Qatar and Russia come online according to the refiner’s schedules.

As can be seen from Figure 3, annual production in 2020, 2021, 2023, and 2024 are all predicted to have a significant annual shortfall compared to predicted demand, perhaps indicating that helium pricing is not set to fall appreciably in the near future.

Figure 4 shows the US Bureau of Land Management auction prices from 2016 to 2019. As can be seen the price of crude helium has risen markedly as the US Bureau of Land Management adjusts auction pricing for crude helium to market values, which is driven up due to the geopolitical factors mentioned above.

Of course, the crude price of helium does not reflect the price paid for bottled helium in the laboratory because further processing and transport costs are incurred by our gas supplier. In fact, the latest spot pricing I have from the UK is a whopping £1011.17 per XL cylinder of grade A helium, which contains 13 cubic metres of helium.

 

So, What is the Truth?

The truth is that there is a lot of information regarding helium supply out there, and much of it is contradictory, misleading, or just plain incorrect. I’ve made a great effort to get facts and figures in this article correct, but if I’ve communicated any misinformation here please let me know.

I now know that, as scientists, we represent a reasonable proportion of the global helium consumption and therefore our voice should be heard regarding the continuity of supply.

I know that the Earth’s reserves will last us for several hundreds of years if consumption does not increase markedly and production continues to be economically viable, with new and expanded production ventures underway in several locations.

However, dwindling reserves in the US and the geopolitical factors associated with helium production and distribution will mean that supply volumes versus consumption will be tight for the next few years, even if planned production increases in Qatar, Russia, and Algeria come online according to schedule. If production from the geothermal spring sources in Tanzania is successful, and depending upon the production capacity achieved, then this shortfall of supply may be alleviated.

Given ongoing uncertainty in the geopolitical landscape, the risk of new production facilities coming onstream according to schedule, and no prospect of helium pricing coming down in the near future, I may just give my gas generator manufacturer a call and start the conversion on my GC and GC–MS instruments. How about you?

References

  1. Incognito, The Column 8(11), 9–10 (2012).
  2. A. Ghosh, Lab Manager11(10), 32–33 (2016).
  3. https://www.peakscientific.com/articles/converting-your-gc-carrier-gas-from-helium-to-hydrogen-:-a-step-by-step-guide/
  4. LCGC North America31(10), 898 (2013).
  5. https://www.petro-online.com/news/analytical-instrumentation/11/bruker-chemical-applied-markets/converting-gc-amp-gc-ms-methods-from-helium-to-hydrogen-carrier-gas-ndash-more-than-just-changingnbspthe-cylinder/28604
  6. https://phys.org/news/2010-08-world-helium-nobel-prize-winner.html
  7. http://www.madehow.com/Volume-4/Helium.html
  8. N.K. Das, H. Chaudhuri, R.K. Bhanderi, et al., Current Science95(12), 1684–1687 (2008).
  9. H. Chaudhuri, B. Sinha, and D. Chandrasekharam, “Helium from Geothermal Sources,” paper presented at the Proceedings World Geothermal Congress 2015, Melbourne, Australia, 2015.
  10. United States Geological Survey, Mineral Commodity Summaries, 2019.
  11. http://www.qatargas.com/english/operations/ras-laffan-helium
  12. The Mineral Industry of Algeria - 2006 Minerals Yearbook, U.S. Geological Survey (April 2008).
  13. http://www.airproducts.com/Company/news-center/2018/11/1114-air-products-and-sonatrach-announce-industrial-gases-agreement-in-algeria.aspx
  14. http://www.gazprominfo.com/terms/orenburg-helium-plant/
  15. https://cen.acs.org/business/instrumentation/Help-helium-users-way/97/i46
  16. https://www.livescience.com/60607-helium-reserve-discovered-in-tanzania-photos.html
  17. https://link.springer.com/article/10.1007%2Fs11053-017-9359-y
  18. Helium – Macro View Update, Edison Investment Research, February 2019.

Contact Author: Incognito
E-mail the Editor:kjones@mmhgroup.com

Recent Videos
Toby Astill | Image Credit: © Thermo Fisher Scientific