Category Archives: Economic and applied geology

Can rock weathering halt global warming?

Read about this at Earth-logs

Maintenance of Earth-pages has stopped. If you wish to continue following reports on significant research developments in Earth science you can register as a follower of my new blog at the Earth-logs site

Ancient oil migration

Read about evidence for petroleum generation and migration through the 1.9 Ga old Gunflint banded iron formation in Ontario, Canada HERE

Interleaved chert (white) and ironstone of the Palaeoproterozoic Gunflint Iron Formation of Ontario, Canada and Minnesota, USA.

UK shale gas: fracking potential dramatically revised downwards

Read about a new estimate of shale-gas reserves in Britain at Earth-logs

Ecological hazards of ocean-floor mining

Read about the threat posed by deep-ocean mining of polymetallic nodules at Earth-logs

ocean floor resources
The distribution of potential ocean-floor metal-rich resources (Credit: Hefferman 2019)

 

Frack me nicely?

Note: Earth-Pages will be closing as of early July, but will continue in another form at Earth-logs

‘There’s a seaside place they call Blackpool that’s famous for fresh air and fun’. Well, maybe, not any more. If you, dear weekender couples, lie still after the ‘fun’ the Earth may yet move for you. Not much, I’ll admit, for British fracking regulations permit Cuadrilla, who have a drill rig at nearby Preston New Road on the Fylde coastal plain of NW England, only to trigger earthquakes with a magnitude less than 0.5 on the Richter scale. This condition was applied after early drilling by Cuadrilla had stimulated earthquakes up to magnitude 3. To the glee of anti-fracking groups the magnitude 0.5 limit has been regularly exceeded, thereby thwarting Cuadrilla’s ambitions from time to time. Leaving aside the view of professional geologists that the pickings for fracked shale gas in Britain [June 2014] are meagre, the methods deployed in hydraulic fracturing of gas-prone shales do pose seismic risks. Geology, beneath the Fylde is about as simple as it gets in tectonically tortured Britain. There are no active faults, and no significant dormant ones near the surface that have moved since about 250 Ma ago; most of Britain is riven by major fault lines, some of which are occasionally active, especially in prospective shale-gas basins near the Pennines. When petroleum companies are bent on fracking they use a drilling technology that allows one site to sink several wells that bend with depth to travel almost horizontally through the target shale rock. A water-based fluid containing a mix of polymers and surfactants to make it slick, plus fine sand or ceramic particles, are pumped at very high pressures into the rock. Joints and bedding in the shale are thus forced open and maintained in that condition by the sandy material, so that gas and even light oil can accumulate and flow up the drill stems to the surface.

Shale, being dominated by ultra-fine clay minerals, is slippery when wet. Consequently, any elastic strain built-up in the rock, either by active tectonics or from long in the past, is likely to be released by fracking. The fractures that release the gas also facilitate the escape of formation water locked in the shale from when it was originally deposited. Being rich in organic matter, target shales maintain highly reducing chemical conditions. So as well as being salty, such formation water may contain high abundances of heavy metals and arsenic, unlike the groundwater in naturally permeable and oxygenated rocks, such as sandstones and limestones. Fracking carries a pollution risk too. Toxic waste fluid is generally disposed of by pumping into permeable strata beneath the well site. There is no knowing where such noxious water might go, other than to follow lines of least resistance, such as large joints and dormant faults that may well be unsuspected at the depths to which drilling might penetrate. That too poses seismic rick by lubrication of the pathways taken by the fluids.

eagle-ford-shale-at-night
The sheer scale of shale-gas fracking in the US is indicated by the light emitted at night by well-lit installations and gas flares in a mature shale-gas basin in Texas targeting the mature, gas-rich Eagle Ford shale. (see: https://geology.com/articles/eagle-ford/)

Britain has barely been touched by fracking or conventional petroleum drilling, unlike large swathes of North America. Fracking began in Kansas, USA in 1947 but got underway in earnest in the 1970s to dominate US natural gas production since the 1990s. The effects of fracking in the long term [July 2013] show up in the active shale-gas basins there. Even in geological settings as quiescent as the Fylde seems to be, the picture is one of repeated earthquakes induced by fracking, which often exceed magnitude 3.0, including one of magnitude 5.6 in Oklahoma that destroyed 14 homes in 2016. A recent paper in Science examines how fluid migration induces dormant structures to move again (Bhattacharya, P. & Viesca, R.C. 2019. Fluid-induced aseismic fault slip outpaces pore-fluid migration. Science, v. 364, p. 464-468; DOI: 10.1126/science.aaw7354). The authors, from Tufts University in the US, used experimental fluid injection in France to indicate that aseismic slip resulting from fluid injection transmits stress far and wide, and more quickly than expected from the outward movement of the injected fluids. This explains why earthquakes produced by deliberate fluid injection into the crust often occur more frequently in active shale-gas basins than they do in areas of naturally high seismic activity

Related article: Fracking: Earthquakes are triggered well beyond fluid injection zones (Science News)

How rich are deep-sea resources?

My first task as a Lecturer in Earth Sciences at the British Open University, from 1971 onward, was to write teaching materials about the economics, formation and geological setting of metal resources. Much of the content was about the full range of ‘conventional’ metal ores, but something being publicised as having huge potential intrigued me. This concerned manganese-rich nodules (with the aesthetic appeal of unwashed potatoes) and crusts found sitting on top of sediments of the abyssal ocean floor, at depths between 3 to 5 kilometres.  While manganese is by no means a rare element and occurs in vast ore reserves on the continents, the nodules contain unusually high concentrations of other, more valuable metals, such as copper, nickel, zinc, cobalt and lead. Some contained more than 3% of Cu, Ni and Co combined, above the ‘grades’ of economic deposits of ores of the individual metals on land. This was the source of their potential: simple, albeit very deep dredging of the nodules would provide multi-metal ore of very high profitability. Moreover, the nodules are in truly vast tonnages (about 10 kg m-2) and continually grow by precipitation from seawater in the underlying sediments at a few millimetres per million years – they are renewable resources.

Manganese nodules taken from the bottom of the...

Manganese nodules from the Pacific abyssal plains. (credit: Wikipedia)

A variety of reasons, not the least of which was the vexatious question of ownership of sea-floor resources far from land, have meant that commercial operations have yet to begin. However, spiralling prices for metals on the world market together with depletion of on-shore, high-grade reserves are beginning to make the opportunity of nodule mining irresistible. Fifteen companies, with licence areas issued by the intergovernmental  International Seabed Authority of around 75 000 km2 each, are now engaged in economic assessment of one of the most remote swathes of the Pacific abyssal plains (Peacock, T. & Alford M.H. 2018. Is deep-sea mining worth it? Scientific American, v. 318(5) (May 2018 issue), p. 63-67). There are several controversial issues surrounding deep-sea mining. First, dredging, like beam trawling disturbs and destroys ocean-floor ecosystems and turns bottom water turbid, the very fine grain size of sediments resulting in settling being very slow ( about 1 mm s-1). Second, preliminary ore processing on board dredging vessels results in plumes of turbid and metal-rich slurry in the wakes, threatening surface and mid-water ecosystems. Such plumes will rapidly spread far from operational areas in surface current systems, eventually to smother pristine areas of ocean floor. Re-examination of areas of experimental dredging from 30 years ago have revealed that they are still sterile of lifeforms larger than 50 micrometres. Added to these effects, onshore processing will produce large amounts of waste – about 75% of the volume of dredged nodules. Conventional mines eventually backfill their excavations, but with nodule mining disposal would be an environmental nightmare.

Japanese sea-floor mining machine. (credit: Japan Times)

Economically, it seems that nodule dredging is potentially highly profitable. To break even requires lifting about a million metric tons, which would yield of the order of 37 000 t of Ni, 32 000 t of Cu, 6000 t of Co and 750 000 t manganese. If all 15 companies begin extraction, production at these levels will have a downward effect on world metal prices, tending to undercut production from conventional mines. One little-considered issue is that the ‘blend’ of metals from nodules will not match the industrial demand for each of them, further destabilising markets. Added to mining of the abyssal plains, plans are well advanced for multi-metal mining of massive sulfide deposits forming at hydrothermal vents or ‘black smokers’ along mid-ocean ridge systems, in which gold figures strongly. Only a few Pacific island states have resisted the ‘promise’ of such operations. Japanese companies are already mining the seabed off Okinawa within their own offshore waters and seemingly are producing zinc equivalent to the country’s annual consumption as well as gold, copper and lead.

‘Big data’ on water resources

 

Two petabytes (2×1015) is a colossal number which happens to approximate how much data has been collected in geocoded form by the Landsat Thematic Mapper and its successors since it was first launched in 1984. In tangible form these would occupy about half a million DVDs, weighing in at about 8 metric tonnes; ‘daunting’ comes nowhere near describing the effort needed to visually interpret this unique set of multi-date imagery. Using the Google Earth Engine, the free cloud-computing platform for big sets of image data which hosts all Landsat data and much else (but not yet the equally daunting ASTER data – roughly a million 136 Mb scenes) the 32 years-worth has been analysed for its content of hydrological information by the European Commission’s Joint Research Centre in Italy, with assistance from Google Switzerland. Using the various spectral characteristics of water in the visible and infrared region, the team has been able to assess the position on the continents of surface water bodies larger than 900 m2, both permanent and ephemeral, and how the various categories have changed in the last 32 years (Pekel, J.-F. et al. 2016. High-resolution mapping of global surface water and its long-term changes. Nature, v. 540, p. 418-422; doi:10.1038/nature20584). The results are conveniently and freely available in their entirety at the Global Surface Water Explorer, an unparalleled and easy-to-use opportunity for water resource managers, wetland ecologists and geographers in general.

Among the revelations are sites and areas that have been subject to gains and losses in water availability, the extents of new and vanished permanent and seasonal water bodies and the conversion of one to the other. A global summary gives a net disappearance of 90 thousand km2 of permanent water bodies, about the area of Lake Superior, but exceeded by new permanent bodies totalling 184 thousand km2. There has been a net increase in permanent water on all continents except Oceania with a loss one percent (note that Antarctica and land north of the Arctic Circle were not analysed). More than 70 % of the losses are in the semi-arid Middle East and Central Asia (Iran, Iraq, Uzbekistan, Kazakhstan and Afghanistan), due mainly to overuse of irrigation, dam construction and long-term drought. Much of the increase in water occurrence stems from reservoir construction, but climate change may have played a part through increased precipitation and melting of high-altitude snow and ice, as in Tibet.

The Aral Sea in Uzbekistan and Kazakhstan has suffered dramatic loss of standing and seasonal water cover due to overuse of water for irrigation from the two main rivers, the Amu (Oxus) and Syr, that flow into it. Note the key to the colours that represent different categories of changes in surface water. (Credit: Global Surface Water Explorer)

The Aral Sea in Uzbekistan and Kazakhstan has suffered dramatic loss of standing and seasonal water cover due to overuse of water for irrigation from the two main rivers, the Amu (Oxus) and Syr, that flow into it. Note the key to the colours that represent different categories of changes in surface water. (Credit: Global Surface Water Explorer)

Many of the lakes in the northern Tibetan Plateau have grown in size during the last 32 years, mainly due to increased precipitation and snow melt. (Credit: Global Surface Water Explorer)

Many of the lakes in the northern Tibetan Plateau have grown in size during the last 32 years, mainly due to increased precipitation and snow melt. (Credit: Global Surface Water Explorer)

There are limitation to the accuracy of the various categories of change, one being the persistence of cloud cover in humid climates, another being the sometimes haphazard scheduling of Landsat Data capture (in some case that has depended on US Government interest in different areas of the world).

More detail on using remote sensing in exploration for and evaluation of water resources can be found here.

Free course on remote sensing for water exploration

250 million people who live in the drylands of Africa and Asia face a shortage of water for their entire lives. Hundreds of millions more in less drought-prone regions of the ‘Third World’ have to cope repeatedly with reduced supplies. A rapid and effective assessment of how to alleviate the shortfall of safe water is therefore vital. In arid and semi-arid areas surface water storage is subject to a greater rate of evaporation than precipitation, so groundwater, hidden beneath the land surface, provides a better alternative. Rainwater is also lost by flowing away far more quickly than in areas with substantial vegetation. Harvesting that otherwise lost resource and diverting it to storage secure from evaporation – ideally by using it to recharge groundwater – is an equally important but less-used strategy. Securing a sustainable water supply for all peoples is the most important objective that geoscientists can address.

In practice, to assure good quality water supplies to a community in the form of productive wells, surface water harvesting schemes or planning the recharge of exploited aquifers requires skill, a great deal of work and considerable financial resources. Yet in many parts of sub-Saharan Africa and arid areas of Asia knowing where to focus effort and increase the chances of it being fruitful is one the biggest hurdles to overcome. Such reconnaissance – highlighting the most probable localities on geological and hydrological grounds, and screening out those least likely to yield water for drinking and hygiene – depends on details of the geology and topography of the terrain in which needy communities are situated. For most of the Afro-Asian dryland belt adequate geological and topographic maps are in as short supply as potable water itself.  Remote sensing combined with an understanding of groundwater storage and surface-water harvesting is a powerful tool for bridging that knowledge gap, and is routinely used successfully in areas blessed with abundances of experienced geoscientists, money and engineering infrastructure. Again, most of the Afro-Asian dryland belt is poorly endowed in these respects.

dvd-sleeve-front

Having long ago written a textbook on general remote sensing for geoscientists, now out of print (Image Interpretation in Geology (3rd edition): 2001. Nelson Thorne/Blackwell Science), I decided to re-issue revised parts of it framed in the specific context of water exploration in arid and semi-arid terrains, and to add practical case studies and exercises based on a free version of professional image processing and desktop mapping software. Some of the most geologically revealing remotely sensed image data – those from the Landsat series of satellites and the joint US-Japan ASTER system carried by Terra, one of NASA’a Earth Observing System satellites – are now easily and freely available for the whole of the Earth’s land surface. Given basic familiarity with theory and practicalities, a computer and appropriate software together with a moderately fast internet connection there is nothing to stop any geoscientist, university geology student or engineer working in the water, sanitation and hygiene (WASH) sector from becoming a proficient, self-contained practitioner in water reconnaissance. Water Exploration: Remote Sensing Approaches has that aim. Online access to the theoretical parts is free, and a DVD that combines theory, software, exemplary data and several exercises that teach the use of image processing/desktop mapping software is available at cost of reproduction and postage.

If you visit the website, find what you see potentially useful and wish to know more, contact me through the Comments form at the H2Oexplore homepage.

Fracking unlikely in Europe

These days, leading British politicians burdened with power have a tendency to show outwardly that they are, if little else, earnest. When busy with economic and industrial policy they wear tailored day-glo hi-viz suits and shiny new hard hats. During the great 2015 floods of Northern England, their garb was off the peg North Face gear and green wellington boots. And, of course, for social policy a hoodie is de rigueur. Rosy-cheeked Prime Minister David Cameron has been extremely earnest about fracking for shale gas for several years, and in the petroleum industry the appropriate signal of a leading politician’s enthusiasm is to wear a rigger’s blue jumpsuit; ‘We’re going all out for shale’ Cameron has said. Given the explosive success of shale-gas exploitation in North America over the last decade that’s not very surprising, but do not expect to see him looking earnestly at an exploration rig again any time soon.

Cameron’s excitement began when in 2011 the Advanced Resources Institute (ARI) in Washington DC released the results of its consultancy for the US Department of Energy on global shale-gas prospects. The star prospect in Europe was Poland, well endowed with subsurface shales, which according to ARI, had more than 5 trillion cubic metres of technically recoverable reserves, enough to satisfy Polish consumption for more than 300 years. In 2013, ARI suggested 17 trillion m3 beneath Britain, albeit only 0.7 trillion that was amenable to fracking (about a decade’s worth of British gas consumption). But still the hype was maintained. An article in the 3 March 2016 issue of Nature (Inman, M. 2016. Can fracking power Europe. Nature, v. 531, p. 22-24) tempers enthusiasm a great deal more.

The Polish Geological Institute revised the country’s reserves down to a tenth of ARI’s estimate. After an initial frenzy of interest following the ARI report, when exploration licences covered a third of Poland, during 2013 and 2014 major companies relinquished licences for fracking en masse. Their exploratory activities had been disappointing because of the depth of burial (2-5 km compared with 1-2 km in the US) and unfavourably high clay content and strength of the target shales. The less thrilling ARI prospects for Britain did not excite major petroleum players at all, what interest there is being from ‘juniors’ such as Cuadrilla. The British Geological Survey, which has huge archives of geological information, both surface and subsurface, has assessed the three main British shale-gas ‘plays’ and comes up with a reserve figure of between 24 and 68 trillion m3. But that high figure is based on the situation in mid-west North American shale-gas fields, where the geology is a great deal simpler than here. In Britain, orogenies at the end of the Carboniferous and the outermost ripples of that which formed the Alps in late Mesozoic and Palaeogene times created far more deformation than beneath the central plains of North America. Widespread faults, even though few in Britain have large displacements, pose two sets of problems. As the minor earthquakes set off by fracking in the tectonically simple Fylde area of western Lancashire indicate, pumping fluids into faulted rock can release pent-up elastic strain. But such leakage into faults and smaller fractures may also cause the injection pressure to fall, making the fracking process less efficient.

https://i0.wp.com/www.agentsofchangefoundation.org/wp-content/uploads/2014/06/4596344953.jpg

Fracking information sheet from the British Geological Survey

Inman reports that fracking is now moribund throughout Europe, partly because of the disappointing results and also because environmental concerns for densely populated regions have spurred widespread moratoria, including those in three of Britain’s four nations; Scotland, Wales and Northern Ireland. The only current European fracking activity is in England, conducted by a handful of junior companies. A stumbling block in England actually lies with the quality of subsurface data for what has been described at the most close examined geology in the world. Since the early 1980s there has been a succession of onshore licensing rounds for oil and conventional gas, the 14th of which is still active. The early ones were accompanied by a great deal of seismic reflection surveying, mainly using the truck mounted ‘Vibroseis’ method where the ground is mechanically thumped rather than subject to explosive shot firing that is favoured in sparsely populated areas. According to BGS, the guardians of the onshore seismic exploration repository, compared with the onshore seismic data available in North America that for Britain is ‘sparse, and fairly poor’.

Paris Agreement 2015: Carbon Capture and Storage

Anyone viewing news that covered the adoption of the Paris Agreement on climate change on 11 December 2015 would have seen clear evidence of the reality of the old saw, ‘There was dancing in the streets’. Not since the premature celebration of the landing of the Philae spacecraft on comet 67P/Churyumov–Gerasimenko 11 months before has there been such public abandonment of normal human restraint. In the case of ‘little Philae’ the object of celebration sputtered out three days after landing, albeit with the collection of some data. Paris 2015 is a great deal more important: the very health of our planet and its biosphere hangs on its successful implementation. At 32 pages long, by UN standards the document agreed to by all 196 UN Member States is pretty succinct considering everything it is supposed to convey to its signatories and the human race at large.

The Bagger 288 bucket wheel reclaimer moves from one lignite mine to another in Germany.

The Bagger 288 bucket wheel reclaimer moves from one lignite mine to another in Germany.

One central and, by most scientific criteria, the most important technology needed as a stopgap before the longed-for adoption of carbon-free energy generation does not figure in the diplomatic screed: carbon capture and storage (CCS) is not mentioned once. Indeed, only 10 Member States have included it in their pledge or ‘intended nationally determined contribution’ (INDC) – Bahrain, Canada, China, Egypt, Iran, Malawi, Norway, Saudi Arabia, South Africa and the United Arab Emirates. Only three of them are notable users of coal-fired power stations for which CCS is most urgent. An article in the January 2016 issue of Scientific American offers an explanation of what seems to be a certain diplomatic timidity about this highly publicized stop-gap measure (Biello, D. 2016. The carbon capture fallacy. Scientific American, v. 314(1) 55-61). David Biello emphasizes the urgency of CCS from more industries than fossil fuel power plants, cement manufacture being a an example. He focuses on the economics and logistics of one of very few CCS facilities that may be on track for commissioning (33 have been shut down or cancelled worldwide since 2010).

The Kemper power station in Mississippi, USA is the most advanced in the US, as it has to be to burn the strip-mined, wet, brown coal or lignite that is its sole fuel. The chemistry it deploys is quite simple but technologically complex and expensive. So Kemper survives only because it aims to sell the captured CO2 to a petroleum company so that it can be pumped into oil fields to increase dwindling production. However, its extraction costs US$1.50 per tonne, while naturally occurring, underground CO2 costs US$0.50 to pump out. Moreover, Kemper’s power output at US$11 000 per kW of generating capacity is three times more expensive than that for a typical coal-fired boiler. Mississippi Power is lucky, in that it only needs to pipe the gas 100 km to its ‘partner’ oil field; a pretty small one producing about 5 000 barrels per day. Some coal plants are near oil fields, but the majority are not. To cap it all, only about a third of the CO2 production is likely to remain in long-term underground storage.

Because Kemper has, predictably, hit the financial buffers (almost US$4 billion over budget) to avoid bankruptcy it has raised electricity prices to its customers by 18%. Without the projected revenue from its partnered oil field it would go belly up. Even in the happy event of financial break-even, in carbon terms it would be subsidising the oilfield to produce…CO2! But the sting in the tail of Biello’s account of this ‘flagship’ project is that the plant is currently neither burning coal nor capturing carbon: it uses natural gas…