share this!

August 29, 2023

This article has been reviewed according to Science X's editorial process and policies . Editors have highlighted the following attributes while ensuring the content's credibility:

fact-checked

peer-reviewed publication

Enhanced chemical weathering: A solution to the climate crisis?

by Petra Giegerich, Johannes Gutenberg-Universität Mainz

Enhanced chemical weathering: A solution to the climate crisis?

Could blending of crushed rock with arable soil lower global temperatures? Researchers of Mainz University have studied global warming events from 40 and 56 million years ago to find answers. Their research paper has recently been published in Nature Geoscience .

The Earth is getting hotter and consequences have been made manifest this summer around the world. Looking back in geological history , global warming events are not uncommon: Around 56 million years ago, during the period known as the Paleocene–Eocene Thermal Maximum (PETM), the temperatures rose by an average of 5 to 8° Celsius. This development was most likely linked to increased volcanism and the associated release of masses of carbon dioxide into the atmosphere. The higher temperatures persisted for about 200,000 years.

Back in 2021, Professor Philip Pogge von Strandmann of Johannes Gutenberg University Mainz (JGU) had already investigated the effect that eventually led to global cooling and climatic recovery after the PETM warming. In short, rainwater combined with the atmospheric carbon dioxide , resulting in carbonic acid that caused enhanced weathering of rock, thus releasing calcium and magnesium.

Rivers then transported the calcium, magnesium, and carbonic acid into the oceans where the calcium, magnesium—and also the carbon dioxide—came together to form insoluble limestone.

"In other words, there is a feedback effect that helps control the climate. High temperatures accelerate the chemical rock weathering process, reducing the levels of carbon dioxide in the atmosphere, allowing the climate to recover," said Pogge von Strandmann.

Climate required twice as long to regenerate 40 million years ago

Climate warming occurred again 16 million years after the PETM during the Middle Eocene Climatic Optimum or MECO. Although volcanic activity resulted in the discharge of roughly the same amounts of carbon dioxide into the atmosphere as during the PETM, it took far longer for the climate to restabilize. The warming effect lasted for an immense 400,000 years, twice as long as in the PETM. Why was recovery so slow during that period?

In searching for an answer, Pogge von Strandmann and co-authors, including first author Alex Krause, began analyzing 40-million-year-old oceanic carbonates and clay minerals to compare the results with those for similar 56-million-year-old examples.

"Just as during the PETM, there was also intensified weathering and erosion in the MECO. However, there was far less exposed rock on the Earth's surface 40 million years ago. Instead, the Earth was extensively covered by a global rainforest the soil of which largely consisted of clay minerals ," explained the researcher.

In contrast with rock, clay does not weather; in fact, it is actually the product of weathering. "So despite the high temperatures, the widespread clay soil prevented rocks from being effectively weathered, a process known as soil shielding," the geoscientist said.

Enhanced weathering for climate protection

How can we use this knowledge in today's world? "We study paleoclimates to determine whether and how we can positively influence our present climate. One option might be to boost the chemical weathering of rock. To help achieve this, we could plow finely crushed rock into our fields," said Pogge von Strandmann. The fine-grained particles of rock would erode rapidly, resulting in the binding of atmospheric carbon dioxide, thus enabling the climate to recuperate.

Negative emissions technologies (NETs) such as this involving the absorption of carbon dioxide are the subjects of intense research across the globe. At the same time, however, if the weathering results in the formation of clay, the effects of the process would be significantly less efficient, as Pogge von Strandmann has discovered. Clay retains the calcium and magnesium that would otherwise be delivered to the ocean. The carbon dioxide would continue to flow into the oceans, but it would not be bound there and would be able to escape back into the atmosphere. In this case, the weathering effect would have next to no influence on the climate.

If the rock particles fully dissolve as a result of weathering, the enhanced weathering concept would turn out to be 100% efficient. However, if all the weathered materials were turned into clay, this would in its turn completely nullify the effect.

In reality, the actual outcome would probably be somewhere between the two extremes: While there was enhanced erosion of rock in the PETM so that the climate normalized more rapidly, clay formation was predominant during the MECO. The extent to which the crushed rock dissolves and how much of it is preserved as clay depends on a range of local factors, such as the globally pre-existing levels of clay and rock.

So in order to establish whether the process of enhanced weathering is a viable approach, it would first be necessary to find out how much clay is formed during the weathering process at each potential location.

Journal information: Nature Geoscience

Provided by Johannes Gutenberg-Universität Mainz

Explore further

Feedback to editors

research about chemical weathering

Extraterrestrial chemistry with earthbound possibilities

2 hours ago

research about chemical weathering

Exoplanets may contain more water than previously thought

research about chemical weathering

Small molecule organic eutectics show potential for replacing plastics

14 hours ago

research about chemical weathering

Animals with higher body temperatures are more likely to evolve into herbivores, study finds

15 hours ago

research about chemical weathering

Study explains why laws are written in an incomprehensible style

16 hours ago

research about chemical weathering

Understanding of early life ecosystems sheds light on evolution of life on Earth

research about chemical weathering

Sea anemone study identifies potentially regenerative stem cells linked to conserved genes

17 hours ago

research about chemical weathering

Ultrathin quantum light sources: Scientists show excitonic interactions boost efficiency of entangled photon generation

18 hours ago

research about chemical weathering

Study finds gun violence in rural America rises as deer hunting season begins

research about chemical weathering

Mapping Martian meteorites: Tracing origins on the red planet

Relevant physicsforums posts, shiveluch volcano erupts on kamchatka peninsula.

Aug 18, 2024

The Secrets of Prof. Verschure's Rosetta Stones

Aug 15, 2024

A very puzzling rock or a pallasite / mesmosiderite or a nothing burger

Aug 14, 2024

M6.8 and M6.3 east of Mindanao, Philippines

Aug 13, 2024

What Could Cause a Persistent 250 Hz Hum at Night?

Aug 12, 2024

M7.1 Earthquake, Hyuganada Sea, E of Kyushu coast

Aug 11, 2024

More from Earth Sciences

Related Stories

research about chemical weathering

How quickly does the climate recover?

Oct 19, 2021

research about chemical weathering

Quantifying carbon leakage from enhanced rock weathering

Apr 5, 2023

research about chemical weathering

North Atlantic volcanic activity was a major driver of climate change 56 million years ago, study finds

Aug 21, 2023

research about chemical weathering

'Planting' rocks in farms, along with emissions reductions, could help meet key IPCC carbon removal goal

Aug 14, 2023

research about chemical weathering

Climate win-win: Study quantifies benefits of enhanced weathering

research about chemical weathering

Assessing weathering conditions around the globe to understand rate-limiting factors for major rock types

Jan 27, 2023

Recommended for you

research about chemical weathering

Scientists investigate triggers of explosive volcanic eruptions in lab simulation study

19 hours ago

research about chemical weathering

High-ambition nitrogen interventions could cut NOₓ emissions by up to 52% by 2050

20 hours ago

research about chemical weathering

Cloud level rise from deforestation threatens water supply in African mountains, study finds

research about chemical weathering

Study explores dynamic role of iron oxides in organic carbon preservation in ocean floor

research about chemical weathering

Energy companies pressure landowners into fracking, study shows

Aug 19, 2024

research about chemical weathering

Geophysicists find link between seismic waves called PKP precursors and strange anomalies in Earth's mantle

Aug 17, 2024

Let us know if there is a problem with our content

Use this form if you have come across a typo, inaccuracy or would like to send an edit request for the content on this page. For general inquiries, please use our contact form . For general feedback, use the public comments section below (please adhere to guidelines ).

Please select the most appropriate category to facilitate processing of your request

Thank you for taking time to provide your feedback to the editors.

Your feedback is important to us. However, we do not guarantee individual replies due to the high volume of messages.

E-mail the story

Your email address is used only to let the recipient know who sent the email. Neither your address nor the recipient's address will be used for any other purpose. The information you enter will appear in your e-mail message and is not retained by Phys.org in any form.

Newsletter sign up

Get weekly and/or daily updates delivered to your inbox. You can unsubscribe at any time and we'll never share your details to third parties.

More information Privacy policy

Donate and enjoy an ad-free experience

We keep our content available to everyone. Consider supporting Science X's mission by getting a premium account.

E-mail newsletter

  • A-Z Publications

Annual Review of Earth and Planetary Sciences

Volume 28, 2000, review article, chemical weathering, atmospheric co 2 , and climate.

  • Lee R. Kump , Susan L. Brantley , and Michael A. Arthur
  • View Affiliations Hide Affiliations Affiliations: 1 Department of Geosciences and Earth System Science Center, The Pennsylvania State University, 302 Deike Building, University Park, Pennsylvania, 16802; email: [email protected]
  • Vol. 28:611-667 (Volume publication date May 2000) https://doi.org/10.1146/annurev.earth.28.1.611
  • © Annual Reviews

There has been considerable controversy concerning the role of chemical weathering in the regulation of the atmospheric partial pressure of carbon dioxide, and thus the strength of the greenhouse effect and global climate. Arguments center on the sensitivity of chemical weathering to climatic factors, especially temperature. Laboratory studies reveal a strong dependence of mineral dissolution on temperature, but the expression of this dependence in the field is often obscured by other environmental factors that co-vary with temperature. In the field, the clearest correlation is between chemical erosion rates and runoff, indicating an important dependence on the intensity of the hydrological cycle. Numerical models and interpretation of the geologic record reveal that chemical weathering has played a substantial role in both maintaining climatic stability over the eons as well as driving climatic swings in response to tectonic and paleogeographic factors.

Article metrics loading...

Full text loading...

Data & Media loading...

  • Article Type: Review Article

Most Read This Month

Most cited most cited rss feed, geologic evolution of the himalayan-tibetan orogen, chemical geodynamics, tectonic implications of the composition of volcanic arc magmas, global glacial isostasy and the surface of the ice-age earth: the ice-5g (vm2) model and grace, oxygen and hydrogen isotopes in the hydrologic cycle, laboratory-derived friction laws and their application to seismic faulting, arc assembly and continental collision in the neoproterozoic east african orogen: implications for the consolidation of gondwanaland, united plates of america, the birth of a craton: early proterozoic assembly and growth of laurentia, biogenic manganese oxides: properties and mechanisms of formation, a tale of amalgamation of three permo-triassic collage systems in central asia: oroclines, sutures, and terminal accretion.

Publication Date: 01 May 2000

Online Option

Sign in to access your institutional or personal subscription or get immediate access to your online copy - available in PDF and ePub formats

  • Press Enter to activate screen reader mode.

Isotope Geochemistry and Cosmochemistry

Chemical weathering.

Chemical weathering is a key Earth surface process, one that shapes the landscape, supplies dissolved constituents to the oceans and, on geological timescales, controls atmospheric CO 2 concentrations.

We study this important process as expressed in river chemistry, in soils, and as recorded in sediments in the oceans and lakes.

research about chemical weathering

What Is Chemical Weathering?

Chemical weathering can change the composition and shape of rocks

DEA PICTURE LIBRARY/Getty Images

  • Types Of Rocks
  • Landforms and Geologic Features
  • Plate Tectonics
  • Weather & Climate
  • B.A., Earth Sciences, University of New Hampshire

There are three types of weathering which affect rock: physical, biological, and chemical. Chemical weathering, also known as decomposition or decay, is the breakdown of rock by chemical mechanisms.

How Chemical Weathering Happens

Chemical weathering does not break rocks into smaller fragments through wind, water, and ice (that's physical weathering ). Nor does it break rocks apart through the action of plants or animals (that's biological weathering). Instead, it changes the chemical composition of the rock, usually through carbonation, hydration, hydrolysis or oxidation. 

Chemical weathering alters the composition of the rock material toward surface minerals , such as clays. It attacks minerals that are relatively unstable in surface conditions, such as the primary minerals of igneous rocks like basalt, granite or peridotite. It can also occur in sedimentary and metamorphic rocks and is an element of  corrosion  or chemical erosion. 

Water is especially effective at introducing chemically active agents by way of fractures and causing rocks to crumble piecemeal. Water may also loosen thin shells of material (in spheroidal weathering). Chemical weathering may include shallow, low-temperature alteration.

Let's take a look at the four main types of chemical weathering that were mentioned earlier. It should be noted that these are not the only forms, just the most common.

Carbonation

Carbonation occurs when rain, which is naturally slightly acidic due to atmospheric carbon dioxide  (CO 2 ), combines with a calcium carbonate (CaCO 3 ), such as limestone or chalk. The interaction forms calcium bicarbonate, or Ca(HCO 3 ) 2 . Rain has a normal pH level of 5.0-5.5, which alone is acidic enough to cause a chemical reaction. Acid rain , which is unnaturally acidic from atmospheric pollution, has a pH level of 4 (a lower number indicates greater acidity while a higher number indicates greater basicity). 

Carbonation, sometimes referred to as dissolution, is the driving force behind the sinkholes, caverns and underground rivers of  karst topography . 

Hydration occurs when water reacts with an anhydrous mineral, creating a new mineral. The water is added to the crystalline structure of a mineral, which forms a hydrate. 

Anhydrite, which means "waterless stone," is a calcium sulfate (CaSO 4 ) that is usually found in underground settings. When exposed to water near the surface, it quickly becomes gypsum , the softest mineral on the Mohs hardness scale .   

Hydrolysis is the opposite of hydration; in this case, water breaks down the chemical bonds of a mineral instead of creating a new mineral. It is a decomposition reaction . 

The name makes this one particularly easy to remember: The prefix "hydro-" means water, while the suffix " -lysis " means decomposition, breakdown or separation. 

Oxidation refers to the reaction of oxygen with metal elements in a rock, forming oxides . An easily recognizable example of this is rust. Iron (steel) reacts easily with oxygen, turning into reddish-brown iron oxides. This reaction is responsible for the red surface of Mars and the red color of hematite and magnetite, two other common oxides.

  • The Definition of Weathering
  • What Is Biological or Organic Weathering of Rocks?
  • Mechanical Weathering
  • Minerals of the Earth's Surface
  • What Is Erosion and How Does It Shape the Earth's Surface?
  • Mechanical Weathering Through Physical Processes
  • What Does Transportation Mean in Geology?
  • What Is Diagenesis in Geology?
  • All About Sediment Grain Size
  • 11 Different Types of Holes in Rocks
  • The Many Variations of Obsidian Rock
  • Breccia Rock Geology and Uses
  • The Basics of Geology
  • The Snowball Earth
  • Marble Rock: Geology, Properties, Uses
  • Geology of the Appalachian Mountains

Weathering describes the breaking down or dissolving of rocks and minerals on the surface of Earth. Water, ice, acids, salts, plants, animals, and changes in temperature are all agents of weathering.

Earth Science, Geology, Geography, Physical Geography

Loading ...

Weathering describes the breaking down or dissolving of rocks and minerals on the surface of Earth. Water, ice, acids , salts , plants, animals, and changes in temperature are all agents of weathering. Once a rock has been broken down, a process called erosion transports the bits of rock and mineral away. No rock on Earth is hard enough to resist the forces of weathering and erosion. Together, these processes carved landmarks such as the Grand Canyon , in the U.S. state of Arizona. This massive canyon is 446 kilometers (277 miles) long, as much as 29 kilometers (18 miles) wide, and 1.6 kilometers (one mile) deep.

Weathering and erosion constantly change the rocky landscape of Earth. Weathering wears away exposed surfaces over time. The length of exposure often contributes to how vulnerable a rock is to weathering. Rocks, such as lavas , that are quickly buried beneath other rocks are less vulnerable to weathering and erosion than rocks that are exposed to agents such as wind and water.

As it smooths rough, sharp rock surfaces, weathering is often the first step in the production of soils . Tiny bits of weathered minerals mix with plants, animal remains , fungi, bacteria, and other organisms. A single type of weathered rock often produces in fertile soil, while weathered materials from a collection of rocks is richer in mineral diversity and contributes to more fertile soil. Soils types associated with a mixture of weathered rock include glacial till , loess , and alluvial sediments .

Weathering is often divided into the processes of mechanical weathering and chemical weathering . Biological weathering , in which living or once-living organisms contribute to weathering, can be a part of both processes.

Mechanical Weathering 

Mechanical weathering, also called physical weathering and disaggregation , causes rocks to crumble. Water, in either liquid or solid form, is often a key agent of mechanical weathering. For instance, liquid water can seep into cracks and crevices in rock. If temperatures drop low enough, the water will freeze . When water freezes, it expands . The ice then works as a wedge . It slowly widens the cracks and splits the rock. When ice melts, liquid water performs the act of erosion by carrying away the tiny rock fragments lost in the split. This specific process (the freeze-thaw cycle) is called frost weathering or cryofracturing .

Temperature changes can also contribute to mechanical weathering in a process called thermal stress . Changes in temperature cause rock to expand (with heat) and contract (with cold). As this happens over and over again, the structure of the rock weakens. Over time, it crumbles. Rocky desert landscapes are particularly vulnerable to thermal stress. The outer layer of desert rocks undergo repeated stress as the temperature changes from day to night. Eventually, outer layers flake off in thin sheets, a process called exfoliation . Exfoliation contributes to the formation of  bornhardts , one of the most dramatic features in landscapes formed by weathering and erosion. Bornhardts are tall, domed, isolated rocks often found in tropical areas. Sugarloaf Mountain, an iconic landmark in Rio de Janeiro, Brazil, is a bornhardt.

Changes in pressure can also contribute to exfoliation due to weathering. In a process called unloading, overlying materials are removed. The underlying rocks, released from overlying pressure, can then expand. As the rock surface expands, it becomes vulnerable to fracturing in a process called sheeting .

Another type of mechanical weathering occurs when clay or other materials near rock absorb water. Clay, more porous than rock, can swell with water, weathering the surrounding, harder rock. Salt also works to weather rock in a process called haloclasty . Saltwater sometimes gets into the cracks and pores of rock. If the saltwater evaporates , salt crystals are left behind. As the crystals grow, they put pressure on the rock, slowly breaking it apart. Honeycomb weathering is associated with haloclasty. As its name implies, honeycomb weathering describes rock formations with hundreds or even thousands of pits formed by the growth of salt crystals. Honeycomb weathering is common in coastal areas, where sea sprays constantly force rocks to interact with salts.

Haloclasty is not limited to coastal landscapes. Salt upwelling , the geologic process in which underground salt domes expand, can contribute to weathering of the overlying rock. Structures in the ancient city of Petra, Jordan, were made unstable and often collapsed due to salt upwelling from the ground below.

Plants and animals can be agents of mechanical weathering. The seed of a tree may sprout in soil that has collected in a cracked rock. As the roots grow, they widen the cracks, eventually breaking the rock into pieces. Over time, trees can break apart even large rocks. Even small plants, such as mosses, can enlarge tiny cracks as they grow. Animals that tunnel underground, such as moles and prairie dogs, also work to break apart rock and soil. Other animals dig and trample rock aboveground, causing rock to slowly crumble. 

Chemical Weathering

Chemical weathering changes the molecular structure of rocks and soil. For instance, carbon dioxide from the air or soil sometimes combines with water in a process called carbonation . This produces a weak acid, called carbonic acid , that can dissolve rock. Carbonic acid is especially effective at dissolving limestone . When carbonic acid seeps through limestone underground, it can open up huge cracks or hollow out vast networks of caves . Carlsbad Caverns National Park, in the U.S. state of New Mexico, includes more than 119 limestone caves created by weathering and erosion. The largest is called the Big Room. With an area of about 33,210 square meters (357,469 square feet), the Big Room is the size of six football fields.

Sometimes, chemical weathering dissolves large portions of limestone or other rock on the surface of Earth to form a landscape called karst . In these areas, the surface rock is pockmarked with holes, sinkholes , and caves. One of the world’s most spectacular examples of karst is Shilin, or the Stone Forest, near Kunming, China. Hundreds of slender, sharp towers of weathered limestone rise from the landscape.

Another type of chemical weathering works on rocks that contain iron. These rocks turn to rust in a process called oxidation . Rust is a compound created by the interaction of oxygen and iron in the presence of water. As rust expands, it weakens rock and helps break it apart.

Hydration is a form of chemical weathering in which the chemical bonds of the mineral are changed as it interacts with water. One instance of hydration occurs as the mineral anhydrite reacts with groundwater . The water transforms anhydrite into gypsum , one of the most common minerals on Earth.

Another familiar form of chemical weathering is hydrolysis . In the process of hydrolysis, a new solution (a mixture of two or more substances) is formed as chemicals in rock interact with water. In many rocks, for example, sodium minerals interact with water to form a saltwater solution.

Hydration and hydrolysis contribute to flared slopes , another dramatic example of a landscape formed by weathering and erosion. Flared slopes are concave rock formations sometimes nicknamed “wave rocks.” Their c-shape is largely a result of subsurface weathering, in which hydration and hydrolysis wear away rocks beneath the landscape’s surface.

Living or once-living organisms can also be agents of chemical weathering. The decaying remains of plants and some fungi form carbonic acid, which can weaken and dissolve rock. Some bacteria can weather rock in order to access nutrients such as magnesium or potassium. Clay minerals, including quartz , are among the most common byproducts of chemical weathering. Clays make up about 40 percent of the chemicals in all sedimentary rocks on Earth.

Weathering and People

Weathering is a natural process, but human activities can speed it up. For example, certain kinds of air pollution increase the rate of weathering. Burning coal , natural gas , and petroleum releases chemicals such as nitrogen oxide and sulfur dioxide into the atmosphere . When these chemicals combine with sunlight and moisture, they change into acids. They then fall back to Earth as acid rain . Acid rain rapidly weathers limestone, marble , and other kinds of stone. The effects of acid rain can often be seen on gravestones , making names and other inscriptions impossible to read. Acid rain has also damaged many historic buildings and monuments . For example, at 71 meters (233 feet) tall, the Leshan Giant Buddha at Mount Emei, China is the world’s largest statue of the Buddha. It was carved 1,300 years ago and sat unharmed for centuries. An innovative drainage system mitigates the natural process of erosion. But in recent years, acid rain has turned the statue’s nose black and made some of its hair crumble and fall.

Spheroidal Weathering

Spheroidal weathering is a form of chemical weathering that occurs when a rectangular block is weathered from three sides at the corners and from two sides along its edges. It is also called “onion skin” weathering.

Weathered Mountains

The Appalachian Mountains in eastern North America once towered more than 9,000 meters (30,000 feet) high—taller than Mount Everest! Over millions of years, weathering and erosion have worn them down. Today, the highest Appalachian peak reaches just 2,037 meters (6,684 feet) high.

Media Credits

The audio, illustrations, photos, and videos are credited beneath the media asset, except for promotional images, which generally link to another page that contains the media credit. The Rights Holder for media is the person or group credited.

Last Updated

April 24, 2024

User Permissions

For information on user permissions, please read our Terms of Service. If you have questions about how to cite anything on our website in your project or classroom presentation, please contact your teacher. They will best know the preferred format. When you reach out to them, you will need the page title, URL, and the date you accessed the resource.

If a media asset is downloadable, a download button appears in the corner of the media viewer. If no button appears, you cannot download or save the media.

Text on this page is printable and can be used according to our Terms of Service .

Interactives

Any interactives on this page can only be played while you are visiting our website. You cannot download interactives.

Related Resources

FSU | Florida State University News

Site Navigation

Global navigation.

Florida State University News

The Official News Source of Florida State University

Home / News / Science & Technology / FSU Research: Chemical weathering could alleviate some climate change effects

FSU Research: Chemical weathering could alleviate some climate change effects

Share on Facebook

There could be some good news on the horizon as scientists try to understand the effects and processes related to climate change.

A team of Florida State University scientists has discovered that chemical weathering, a process in which carbon dioxide breaks down rocks and then gets trapped in sediment, can happen at a much faster rate than scientists previously assumed and could potentially counteract some of the current and future climate change caused by humans.

The findings were published in the journal Scientific Reports .

Scientists have generally thought that this process takes hundreds of thousands to millions of years to occur, helping to alleviate warming trends at an exceptionally slow rate.

Rather than potentially millions of years, FSU researchers now suggest it can take several tens of thousands of years.

It’s not a quick fix though.

“Increased chemical weathering is one of Earth’s natural responses to carbon dioxide increases,” said Theodore Them, the lead researcher on the paper and a postdoctoral researcher at Florida State and the National High Magnetic Field Laboratory. “The good news is that this process can help balance the effects of fossil fuel combustion, deforestation and agricultural practices. The bad news is that it will not begin to counteract the excessive amounts of atmospheric carbon dioxide that humans are emitting for at least several thousand years.”

As atmospheric carbon dioxide concentrations increase, the climate gets warmer. The warmer climate speeds up chemical weathering, which consumes carbon dioxide from the atmosphere and mitigates the greenhouse effect, thus leading to climate cooling.

To conduct the study, the research team determined the rate at which rocks were chemically broken down over a period of rapid warming in the Early Jurassic Epoch called the Toarcian Oceanic Anoxic Event, an interval where a major extinction event occurred about 183 million years ago.

Working with colleagues at Durham University in the United Kingdom and using state-of-the-art analytical instrumentation within the National MagLab’s  Geochemistry Group , the researchers processed and measured the trace elements of their rock samples.

“We noticed that although chemical weathering increased significantly during this time interval, it was not as large as previously hypothesized for this event,” Them said. “What’s really striking, however, is the planet’s ability to respond to these environmental changes on such short timescales.”

This increased chemical weathering process could have another downside.

The researchers’ findings suggest that widespread oxygen-deficient oceans occurred because an excess of nutrients from the breakdown of rocks flowed into the oceans during the Early Jurassic Period.

The researchers predict that future changes in climate and weather patterns due to a warming planet will create more precipitation and increase the amount of river water and nutrients transported to coastal regions. This is expected to increase both the size and duration of future coastal ocean deoxygenation, negatively impacting sea life in those areas.

“Understanding ancient climatic change like this helps us anticipate the timing, implications, and environmental response to better predict future climate scenarios,” said FSU Assistant Professor of Geology Jeremy Owens, a co-author on the paper.

Other authors on the paper include Benjamin Gill from Virginia Tech, David Selby and Darren Gröcke from Durham University and Richard Friedman from University of British Columbia.

Resolving controversies surrounding carbon sinks from carbonate weathering

  • Published: 08 August 2024
  • Volume 67 , pages 2705–2717, ( 2024 )

Cite this article

research about chemical weathering

  • Xiaoyong Bai 1 , 4 , 6 ,
  • Sirui Zhang 1 , 2 ,
  • Pete Smith 3 ,
  • Chaojun Li 1 , 2 ,
  • Lian Xiong 1 , 5 ,
  • Chaochao Du 1 , 5 ,
  • Yingying Xue 1 , 5 ,
  • Zilin Li 1 , 5 ,
  • Mingkang Long 1 , 2 ,
  • Minghui Li 1 , 5 ,
  • Xiaoyun Zhang 1 , 5 ,
  • Shu Yang 1 , 5 ,
  • Qing Luo 1 , 5 &
  • Xiaoqian Shen 1 , 5  

Explore all metrics

The importance of carbonate weathering carbon sinks (CCSs) is almost equal to that of vegetation photosynthesis in the global carbon cycle. However, CCSs have become controversial in formulating carbon neutral policies to deal with global climate problems in various countries, since the carbonate dissolution is reversible. In order to address these controversies, we reviewed recent advances in understanding CCSs and examined the outstanding controversies surrounding them. We have analyzed the five controversies, revealing the existence of CCSs, quantifying their magnitude, clarifying their spatiotemporal pattern, and documenting how they have increased and how they evolved under the background of global change. By addressing these five controversies, we help to bring clarity to the role of CCSs in the carbon cycle of global terrestrial ecosystems.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

Atjay G L, Ketner P, Duvigneaud P. 1979. Terrestrial primary production and phytomass. In: Bolin B, Degens E T, Kempe S, Ketner P, eds. The Global Carbon Cycle. New York: John Wiley & Sons. 129–181.

Google Scholar  

Allegre C J, Schneider S H. 1994. The evolution of the Earth. Sci Am, 271: 66–75

Article   CAS   Google Scholar  

Amiotte Suchet P, Probst J L, Ludwig W. 2003. Worldwide distribution of continental rock lithology: Implications for the atmospheric/soil CO 2 uptake by continental weathering and alkalinity river transport to the oceans. Glob Biogeochem Cycle, 17: 1038

Article   Google Scholar  

Anderson L G, Kaltin S. 2001. Carbon fluxes in the Arctic Ocean—Potential impact by climate change. Polar Res, 20: 225–232

Andrews J A, Schlesinger W H. 2001. Soil CO 2 dynamics, acidification, and chemical weathering in a temperate forest with experimental CO 2 enrichment. Glob Biogeochem Cycle, 15: 149–162

Battin T J, Luyssaert S, Kaplan L A, Aufdenkampe A K, Richter A, Tranvik L J. 2009. The boundless carbon cycle. Nat Geosci, 2: 598–600

Beaulieu E, Goddéris Y, Donnadieu Y, Labat D, Roelandt C. 2012. High sensitivity of the continental-weathering carbon dioxide sink to future climate change. Nat Clim Change, 2: 346–349

Berner R A. 1997. The rise of plants and their effect on weathering and atmospheric CO 2 . Science, 276: 544–546

Berner R A. 2006. GEOCARBSULF: A combined model for Phanerozoic atmospheric O 2 and CO 2 . Geochim Cosmochim Acta, 70: 5653–5664

Berner E K, Berner R A. 1987. The Global Water Cycle—Geochemistry and Environment. Englewood Cliffs: Prentice-Hall

Bingeman C W, Varner J E, Martin W P. 1953. The effect of the addition of organic materials on the decomposition of an organic soil. Soil Sci Soc Amer J, 17: 34–38

Bufe A, Hovius N, Emberson R, Rugenstein J K C, Galy A, Hassenruck-Gudipati H J, Chang J M. 2021. Co-variation of silicate, carbonate and sulfide weathering drives CO 2 release with erosion. Nat Geosci, 14: 211–216

Burke A, Present T M, Paris G, Rae E C M, Sandilands B H, Gaillardet J, Peucker-Ehrenbrink B, Fischer W W, McClelland J W, Spencer R G M, Voss B M, Adkins J F. 2018. Sulfur isotopes in rivers: Insights into global weathering budgets, pyrite oxidation, and the modern sulfur cycle. Earth Planet Sci Lett, 496: 168–177

Cao J H, Jiang Z C, Yuan D X, Xia R Y, Zhang C. 2017. The progress in the study of the karst dynamic system and global changes in the past 30 years. Geol China, 44: 874–900

CAS   Google Scholar  

Carter A J, Scholes R J. 2000. Spatial Global Database of Soil Properties. GBP Global Soil Data Task. International Toulouse: Geosphere-Biosphere Programme (IGBP) Data Information Systems

Carvalhais N, Forkel M, Khomik M, Bellarby J, Jung M, Migliavacca M, Mu M, Saatchi S, Santoro M, Thurner M, Weber U, Ahrens B, Beer C, Cescatti A, Randerson J T, Reichstein M. 2014. Global covariation of carbon turnover times with climate in terrestrial ecosystems. Nature, 514: 213–217

Chang J F, Ciais P, Gasser T, Smith P, Herrero M, Havlík P, Obersteiner M, Guenet B, Goll D S, Li W, Naipal V, Peng S S, Qiu C J, Tian H Q, Viovy N, Yue C, Zhu D. 2021. Climate warming from managed grasslands cancels the cooling effect of carbon sinks in sparsely grazed and natural grasslands. Nat Commun, 12: 118

Chen C C, Gong G C, Shiah F K. 2007. Hypoxia in the East China Sea: One of the largest coastal low-oxygen areas in the world. Mar Environ Res, 64: 399–408

Chen H, Ju P J, Zhu Q, Xu X L, Wu N, Gao Y H, Feng X J, Tian J Q, Niu S L, Zhang Y J, Peng C H, Wang Y F. 2022. Carbon and nitrogen cycling on the Qinghai-Tibetan Plateau. Nat Rev Earth Environ, 3: 701–716

Ciais P, Sabine C, Bala G, Bopp L, Brovkin V, Canadell J, Chhabra A, Defries R, Galloway J, Heimann M, Jones C, Quéré C L, Myneni R B, Piao S, Thornton P. 2013. Carbon and Other Biogeochemical Cycles. Cambridge: Cambridge University Press

Cole J J, Prairie Y T, Caraco N F, McDowell W H, Tranvik L J, Striegl R G, Duarte C M, Kortelainen P, Downing J A, Middelburg J J, Melack J. 2007. Plumbing the global carbon cycle: Integrating inland waters into the terrestrial carbon budget. Ecosystems, 10: 172–185

Copard Y, Amiotte-Suchet P, Di-Giovanni C. 2007. Storage and release of fossil organic carbon related to weathering of sedimentary rocks. Earth Planet Sci Lett, 258: 345–357

Cox P M, Pearson D, Booth B B, Friedlingstein P, Huntingford C, Jones C D, Luke C M. 2013. Sensitivity of tropical carbon to climate change constrained by carbon dioxide variability. Nature, 494: 341–344

Curl R L. 2012. Carbon shifted but not sequestered. Science, 335: 655

de la Puente L, Pedro Ferrio J, Palacio S. 2022. Disentangling water sources in a gypsum plant community. Gypsum crystallization water is a key source of water for shallow-rooted plants. Ann Bot, 129: 87–100

Drake T W, Tank S E, Zhulidov A V, Holmes R M, Gurtovaya T, Spencer R G M. 2018. Increasing alkalinity export from large Russian Arctic rivers. Environ Sci Technol, 52: 8302–8308

Dreybrodt W. 2012. Processes in Karst Systems: Physics, Chemistry, and Geology (Vol. 4). Cham: Springer Science & Business Media

Du C, Bai X, Li Y, Tan Q, Zhao C, Luo G, Wu L, Chen F, Li C, Ran C, Zhang S, Xiong L, Song F, Xiao B, Li Z, Xue Y, Long M, Luo Q, Zhang X, Li M, Shen X, Yang S. 2024. Storage, form, and influencing factors of karst inorganic carbon in a carbonate area in China. Sci China Earth Sci, 67: 725–739

Erb K H, Kastner T, Plutzar C, Bais A L S, Carvalhais N, Fetzel T, Gingrich S, Haberl H, Lauk C, Niedertscheider M, Pongratz J, Thurner M, Luyssaert S. 2018. Unexpectedly large impact of forest management and grazing on global vegetation biomass. Nature, 553: 73–76

Falkowski P, Scholes R J, Boyle E, Canadell J, Canfield D, Elser J, Gruber N, Hibbard K, Högberg P, Linder S, Mackenzie F T, Moore III B, Pedersen T, Rosenthal Y, Seitzinger S, Smetacek V, Steffen W. 2000. The global carbon cycle: A test of our knowledge of earth as a system. Science, 290: 291–296

Fan B L, Zhao Z Q, Tao F X, Liu B J, Tao Z H, Gao S, Zhang L H. 2014. Characteristics of carbonate, evaporite and silicate weathering in Huanghe River basin: A comparison among the upstream, midstream and downstream. J Asian Earth Sci, 96: 17–26

Ferris F G, Wiese R G, Fyfe W S. 1994. Precipitation of carbonate minerals by microorganisms: Implications for silicate weathering and the global carbon dioxide budget. Geomicrobiol J, 12: 1–13

Foley J A. 1995. An equilibrium model of the terrestrial carbon budget. Tellus B-Chem Phys Meteor, 47: 310–319

Ford D, William P. 2007. Karst hydrogeology and geomorphology. Chichester: John Wiley & Sons Ltd. 1–562

Book   Google Scholar  

Friedlingstein P, O’Sullivan M, Jones M W, Andrew R M, Hauck J, Olsen A, Peters G P, Peters W, Pongratz J, Sitch S, Le Quéré C, Canadell J G, Ciais P, Jackson R B, Alin S, Aragão L E O C, Arneth A, Arora V, Bates N R, Becker M, Benoit-Cattin A, Bittig H C, Bopp L, Bultan S, Chandra N, Chevallier F, Chini L P, Evans W, Florentie L, Forster P M, Gasser T, Gehlen M, Gilfillan D, Gkritzalis T, Gregor L, Gruber N, Harris I, Hartung K, Haverd V, Houghton R A, Ilyina T, Jain A K, Joetzjer E, Kadono K, Kato E, Kitidis V, Korsbakken J I, Landschützer P, Lefèvre N, Lenton A, Lienert S, Liu Z, Lombardozzi D, Marland G, Metzl N, Munro D R, Nabel J E M S, Nakaoka S I, Niwa Y, O’Brien K, Ono T, Palmer P I, Pierrot D, Poulter B, Resplandy L, Robertson E, Rödenbeck C, Schwinger J, Séférian R, Skjelvan I, Smith A J P, Sutton A J, Tanhua T, Tans P P, Tian H, Tilbrook B, van der Werf G, Vuichard N, Walker A P, Wanninkhof R, Watson A J, Willis D, Wiltshire A J, Yuan W, Yue X, Zaehle S. 2020. Global carbon budget 2020. Earth Syst Sci Data, 12: 3269–3340

Gaillardet J, Galy A. 2008. Himalaya—Carbon sink or source? Science, 320: 1727–1728

Gaillardet J, Dupré B, Louvat P, Allègre C J. 1999. Global silicate weathering and CO 2 consumption rates deduced from the chemistry of large rivers. Chem Geol, 159: 3–30

Gaillardet J, Calmels D, Romero-Mujalli G, Zakharova E, Hartmann J. 2019. Global climate control on carbonate weathering intensity. Chem Geol, 527: 118762

Gatti L V, Gloor M, Miller J B, Doughty C E, Malhi Y, Domingues L G, Basso L S, Martinewski A, Correia C S C, Borges V F, Freitas S, Braz R, Anderson L O, Rocha H, Grace J, Phillips O L, Lloyd J. 2014. Drought sensitivity of Amazonian carbon balance revealed by atmospheric measurements. Nature, 506: 76–80

George E F, Hall M A, Klerk G. 2008. Plant Propagation by Tissue Culture. Amsterdam: Springer Netherlands. 283–333

Gislason S R, Oelkers E H, Eiriksdottir E S, Kardjilov M I, Gisladottir G, Sigfusson B, Snorrason A, Elefsen S, Hardardottir J, Torssander P, Oskarsson N. 2009. Direct evidence of the feedback between climate and weathering. Earth Planet Sci Lett, 277: 213–222

Goldscheider N, Chen Z, Auler A S, Bakalowicz M, Broda S, Drew D, Hartmann J, Jiang G H, Moosdorf N, Stevanovic Z, Veni G. 2020. Global distribution of carbonate rocks and karst water resources. Hydrogeol J, 28: 1661–1677

Goll D S, Ciais P, Amann T, Buermann W, Chang J F, Eker S, Hartmann J, Janssens I, Li W, Obersteiner M, Penuelas J, Tanaka K, Vicca S. 2021. Potential CO 2 removal from enhanced weathering by ecosystem responses to powdered rock. Nat Geosci, 14: 545–549

Gombert P. 2002. Role of karstic dissolution in global carbon cycle. Glob Planet Change, 33: 177–184

Gong S H, Wang S J, Bai X Y, Luo G J, Wu L H, Chen F, Qian Q H, Xiao J Y, Zeng C. 2021. Response of the weathering carbon sink in terrestrial rocks to climate variables and ecological restoration in China. Sci Total Environ, 750: 141525

Groves C, Cao J H, Zhang C. 2012. Response-carbon shifted but not sequestered. Science, 335: 655

Hartmann J. 2009. Bicarbonate-fluxes and CO 2 -consumption by chemical weathering on the Japanese Archipelago—Application of a multi-lithological model framework. Chem Geol, 265: 237–271

Hartmann J, Moosdorf N, Lauerwald R, Hinderer M, West A J. 2014. Global chemical weathering and associated P-release—The role of lithology, temperature and soil properties. Chem Geol, 363: 145–163

Hengl T, de Jesus J M, MacMillan R A, Batjes N H, Heuvelink G B M, Ribeiro E, Samuel-Rosa A, Kempen B, Leenaars J G B, Walsh M G, Gonzalez M R. 2014. SoilGrids1km—Global soil information based on automated mapping. PLoS ONE, 9: e105992

Hilley G E, Porder S. 2008. A framework for predicting global silicate weathering and CO 2 drawdown rates over geologic time-scales. Proc Natl Acad Sci USA, 105: 16855–16859

Hilton R G, West A J. 2020. Mountains, erosion and the carbon cycle. Nat Rev Earth Environ, 1: 284–299

Houghton R A, Hobbie J E, Melillo J M, Moore B, Peterson B J, Shaver G R, Woodwell G M. 1983. Changes in the carbon content of terrestrial biota and soils between 1860 and 1980: A net release of CO 2 to the atmosphere. Ecol Monogr, 53: 235–262

Huang F, Wu X, Yang H, Zhang C L, Cao J H. 2018. Study on carbon cycle of karst critical zone in Maocun subterranean river basin of Guilin. Guangxi Sci, 25: 515–523

Huang M T, Piao S L, Ciais P, Peñuelas J, Wang X H, Keenan T F, Peng S S, Berry J A, Wang K, Mao J F, Alkama R, Cescatti A, Cuntz M, Deurwaerder DH, Gao M D, He Y, Liu Y W, Luo Y Q, Myneni R B, Niu S L, Shi X Y, Yuan W Q, Verbeeck H, Wang T, Wu J, Janssens I A. 2019. Air temperature optima of vegetation productivity across global biomes. Nat Ecol Evol, 3: 772–779

IPCC. 2013. Climate Change 2013: The Physical Science Basis. In: Stocker T F, Qin D, Plattner G K, Tignor M, Allen S K, Boschung J, Nauels A, Xia Y, Bex V, Midgley P M, eds. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press. 1535

Jain A K, West T O, Yang X J, Post W M. 2005. Assessing the impact of changes in climate and CO 2 on potential carbon sequestration in agricultural soils. Geophys Res Lett, 32: L19711

Jiang Z C, Qin X Q, Cao J H, He S Y, Zhang C, Zhang Q. 2013. Significance and carbon sink effects of karst processes in global carbon cycle—Also reply to Discussion on article “ Calculation of atmospheric CO 2 sink formed in karst processes of karst divided regions in China ” (in Chinese). Carsol Sin, 32: 1–6

Jiang Z C, Zhang C, Luo W Q, Xiao Q, Wu Z Y. 2022. Research progress and prospect of carbon sink in karst region of China (in Chinese). Carsol Sin, 41: 345–355

Jobbágy E G, Jackson R B. 2000. The vertical distribution of soil organic carbon and its relation to climate and vegetation. Ecol Appl, 10: 423–436

Kasting J F. 1987. Theoretical constraints on oxygen and carbon dioxide concentrations in the Precambrian atmosphere. Precambrian Res, 34: 205–229

Khadka M B, Martin J B, Jin J. 2014. Transport of dissolved carbon and CO 2 degassing from a river system in a mixed silicate and carbonate catchment. J Hydrol, 513: 391–402

Köhler P. 2020. Anthropogenic CO 2 of high emission scenario compensated after 3500 years of ocean alkalinization with an annually constant dissolution of 5 Pg of olivine. Front Clim, 2: 575744

King A W, Post W M, Wullschleger S D. 1997. The potential response of terrestrial carbon storage to changes in climate and atmospheric CO 2 . Clim Change, 35: 199–227

Komar N, Zeebe R E. 2021. Reconciling atmospheric CO 2 , weathering, and calcite compensation depth across the Cenozoic. Sci Adv, 7: eabd4876

Lal R. 2004. Soil carbon sequestration to mitigate climate change. Geoderma, 123: 1–22

Lapenis A, Shvidenko A, Shepaschenko D, Nilsson S, Aiyyer A. 2005. Acclimation of Russian forests to recent changes in climate. Glob Change Biol, 11: 2090–2102

Larson C. 2011. An unsung carbon sink. Science, 334: 886–887

Li C J, Bai X Y, Tan Q, Luo G J, Wu L H, Chen F, Xi H P, Luo X L, Ran C, Chen H, Zhang S R, Liu M, Gong S H, Xiong L, Song F, Xiao B, Du C C. 2022a. High-resolution mapping of the global silicate weathering carbon sink and its long-term changes. Glob Change Biol, 28: 4377–4394

Li H W, Wang S J, Bai X Y, Luo W J, Tang H, Cao Y, Wu L H, Chen F, Li Q, Zeng C, Wang M M. 2018. Spatiotemporal distribution and national measurement of the global carbonate carbon sink. Sci Total Environ, 643: 157–170

Li H W, Wang S J, Bai X Y, Cao Y, Wu L H. 2019a. Spatiotemporal evolution of carbon sequestration of limestone weathering in China. Sci China Earth Sci, 62: 974–991

Li H W, Wang S J, Bai X Y, Cao Y, Tian Y C, Luo G J, Chen F, Li Q, Wu L H, Wang J F, Wang M M, Tian S Q, Deng Y H, Hu Z Y, Yang Y J, Li C J, Lu Q, Xi H P, Chen H, Ran C, Luo X L. 2019b. Effects of climate change and ecological restoration on carbonate rock weathering carbon sequestration in the karst valley of Southwest China (in Chinese). Acta Ecol Sin, 39: 6158–6172

Li J Y, Zhang J. 2003. Chemical weathering processes and atmospheric CO 2 consumption in the Yellow River drainage basin (in Chinese). Marine Geol Quat Geol, 2: 43–49

Li L, Zhang Y, Zhou T, Wang K, Wang C, Wang T, Yuan L, An K, Zhou C, Lü G. 2022b. Mitigation of China’s carbon neutrality to global warming. Nat Commun, 13: 5315

Li Y, Zhang C Q, Wang N A, Han Q, Zhang X Z, Liu Y, Xu L M, Ye W T. 2017. Substantial inorganic carbon sink in closed drainage basins globally. Nat Geosci, 10: 501–506

Liu J K, Han G L. 2020. Major ions and δ 34 S SO4 in Jiulongjiang River water: Investigating the relationships between natural chemical weathering and human perturbations. Sci Total Environ, 724: 138208

Liu L, Zhuang Q L, Zhao D S, Zheng D, Kou D, Yang Y H. 2022. Permafrost degradation diminishes terrestrial ecosystem carbon sequestration capacity on the Qinghai-Tibetan Plateau. Glob Biogeochem Cycle, 36: e2021GB007068

Liu Z H, Zhao J. 2000. Contribution of carbonate rock weathering to the atmospheric CO 2 sink. Environ Geol, 39: 1053–1058

Liu Z H, Dreybrodt W. 2012. Comparison of carbon sequestration capacity between carbonate weathering and forests: The necessity to change traditional ideas and methods of study of carbon sinks (in Chinese). Carsol Sin, 31: 345–348

Liu Z H, Dreybrodt W. 2015. Significance of the carbon sink produced by H 2 O-carbonate-CO 2 -aquatic phototroph interaction on land. Sci Bull, 60: 182–191

Liu Z H, Dreybrodt W, Wang H J. 2010. A new direction in effective accounting for the atmospheric CO 2 budget: Considering the combined action of carbonate dissolution, the global water cycle and photosynthetic uptake of DIC by aquatic organisms. Earth-Sci Rev, 99: 162–172

Liu Z H, Dreybrodt W, Liu H. 2011. Atmospheric CO 2 sink: Silicate weathering or carbonate weathering? Appl Geochem, 26: S292–S294

Liu Z H, Macpherson G L, Groves C, Martin J B, Yuan D X, Zeng S B. 2018. Large and active CO 2 uptake by coupled carbonate weathering. Earth-Sci Rev, 182: 42–49

Liu Z H, Yan H, Zeng S B. 2021. Increasing autochthonous production in inland waters as a contributor to the missing carbon sink. Front Earth Sci, 9: 620513

Luo W Q, Jiang Z C, Zhao C, Yang F J, Deng Y. 2009. Study on calcareous soil amelioration with filtered sludge of sugar refinery (in Chinese). Res Agricult Moderniz, 30: 248–252

Martin J B. 2017. Carbonate minerals in the global carbon cycle. Chem Geol, 449: 58–72

Merkel B, Planer F B. 2005. Groundwater Geochemistry. Berlin Heideberg: Springer-Verlag. 1–200

Meybeck M. 1987. Global chemical weathering of surficial rocks estimated from river dissolved loads. Am J Sci, 287: 401–428

Meyers P A, Ishiwatari R. 1993. Lacustrine organic geochemistry—An overview of indicators of organic matter sources and diagenesis in lake sediments. Org Geochem, 20: 867–900

Mitsch W J, Bernal B, Nahlik A M, Mander Ü, Zhang L, Anderson C J, Jørgensen S E, Brix H. 2013. Wetlands, carbon, and climate change. Landscape Ecol, 28: 583–597

Moon S, Chamberlain C P, Hilley G E. 2014. New estimates of silicate weathering rates and their uncertainties in global rivers. Geochim Cosmochim Acta, 134: 257–274

Morse J W, Arvidson R S, Lüttge A. 2007. Calcium carbonate formation and dissolution. Chem Rev, 107: 342–381

Mumby P J, Chollett I, Bozec Y M, Wolff N H. 2014. Ecological resilience, robustness and vulnerability: How do these concepts benefit ecosystem management? Curr Opin Environ Sust, 7: 22–27

Nguyen A T Q, Nguyen A M, Pham N T T, Nguyen H X, Dang Q T, Tran T M, Nguyen A D, Tran P D, Nguyen M N. 2021. CO 2 can decrease the dissolution rate of ashed phytoliths. Geoderma, 385: 114835

Olson J S, Watts J A, Allison L J. 1983. Carbon in live vegetation of major world ecosystems. Oak Ridge National Laboratory

Pan Y, Birdsey R A, Fang J, Houghton R, Kauppi P E, Kurz W A, Phillips O L, Shvidenko A, Lewis S L, Canadell J G, Ciais P, Jackson R B, Pacala S W, McGuire A D, Piao S, Rautiainen A, Sitch S, Hayes D. 2011. A large and persistent carbon sink in the world’s forests. Science, 333: 988–993

Porder S. 2019. How plants enhance weathering and how weathering is important to plants. Elements, 15: 241–246

Post W M, Emanuel W R, Zinke P J, Stangenberger A G. 1982. Soil carbon pools and world life zones. Nature, 298: 156–159

Poulter B, Frank D, Ciais P, Myneni R B, Andela N, Bi J, Broquet G, Canadell J G, Chevallier F, Liu Y Y, Running S W, Sitch S, van der Werf G R. 2014. Contribution of semi-arid ecosystems to interannual variability of the global carbon cycle. Nature, 509: 600–603

Prentice I C, Sykes M T, Lautenschlager M, Harrison S P, Denissenko O, Bartlein P J. 1993. Modelling global vegetation patterns and terrestrial carbon storage at the Last Glacial Maximum. Glob Ecol Biogeography Lett, 3: 67–76

Prentice I C, Farquhar G, Fasham M J R, Houghton J T, Ding Y, Griggs D J. 2001. The Science of Climate Change. Contributions of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change.Climate Change, 183–237

Pu J, Li J, Zhang T, Xiong X, Yuan D. 2019. High spatial and seasonal heterogeneity of p CO 2 and CO 2 emissions in a karst groundwater-stream continuum, southern China. Environ Sci Pollut Res, 26: 25733–25748

Pugh T A M, Lindeskog M, Smith B, Poulter B, Arneth A, Haverd V, Calle L. 2019. Role of forest regrowth in global carbon sink dynamics. Proc Natl Acad Sci USA, 116: 4382–4387

Qin X Q, Liu P, Huang Q B, Zhang L K. 2013. Estimation of atmospheric/soil CO 2 consumption by rock weathering in the Pearl River Valley (in Chinese). Acta Geosci Sin, 34: 455–462

Raymond P A, Oh N H, Turner R E, Broussard W. 2008. Anthropogenically enhanced fluxes of water and carbon from the Mississippi River. Nature, 451: 449–452

Reay D S, Dentener F, Smith P, Grace J, Feely R A. 2008. Global nitrogen deposition and carbon sinks. Nat Geosci, 1: 430–437

Ren W, Banger K, Tao B, Yang J, Huang Y W, Tian H Q. 2020. Global pattern and change of cropland soil organic carbon during 1901-2010: Roles of climate, atmospheric chemistry, land use and management. Geogr Sustain, 1: 59–69

Romero-Mujalli G, Hartmann J, Börker J. 2018a. Temperature and CO 2 dependency of global carbonate weathering fluxes—Implications for future carbonate weathering research. Chem Geol, 527: 118874

Romero-Mujalli G, Hartmann J, Börker J, Gaillardet J, Calmels D. 2018b. Ecosystem controlled soil-rock p CO 2 and carbonate weathering—Constraints by temperature and soil water content. Chem Geol, 527: 118634

Schindlbacher A, Beck K, Holzheu S, Borken W. 2019. Inorganic carbon leaching from a warmed and irrigated carbonate forest soil. Front For Glob Change, 2: 40

Sha Z Y, Bai Y F, Li R R, Lan H, Zhang X L, Li J, Liu X F, Chang S J, Xie Y C. 2022. The global carbon sink potential of terrestrial vegetation can be increased substantially by optimal land management. Commun Earth Environ, 3: 8

Shan X, Liu P, Zhang L, Qin X, Cao J, Yu D, Yao X, Liu W. 2019. Hydrochemical characteristics and karst carbon sink estimation under the influence of allogenic water. Carbonates Evaporites, 34: 1855–1863

Singer M J, Warkentin B P. 1996. Soils in an environmental context: An American perspective. Catena, 27: 179–189

Sipos J, Hodecek J, Kuras T, Dolny A. 2017. Principal determinants of species and functional diversity of carabid beetle assemblages during succession at post-industrial sites. Bull Entomol Res, 107: 466–477

Soulet G, Hilton R G, Garnett M H, Roylands T, Klotz S, Croissant T, Dellinger M, Le Bouteiller C. 2021. Temperature control on CO 2 emissions from the weathering of sedimentary rocks. Nat Geosci, 14: 665–671

Spence J, Telmer K. 2005. The role of sulfur in chemical weathering and atmospheric CO 2 fluxes: Evidence from major ions, δ 13 C DIC , and δ 34 S SO4 in rivers of the Canadian Cordillera. Geochim Cosmochim Acta, 69: 5441–5458

Strefler J, Amann T, Bauer N, Kriegler E, Hartmann J. 2018. Potential and costs of carbon dioxide removal by enhanced weathering of rocks. Environ Res Lett, 13: 034010

Subramaniam A, Yager P L, Carpenter E J, Mahaffey C, Björkman K, Cooley S, Kustka A B, Montoya J P, Sañudo-Wilhelmy S A, Shipe R, Capone D G. 2008. Amazon River enhances diazotrophy and carbon sequestration in the tropical North Atlantic Ocean. Proc Natl Acad Sci USA, 105: 10460–10465

Sulla-Menashe D, Friedl M A. 2018. User guide to collection 6 MODIS land cover (MCD12Q1 and MCD12C1) product. Reston, VA, USA: USGS. 18

Task G S D. 2014. Global soil data products CD-ROM contents (IGBPDIS). ORNL DAAC

Tian H Q, Melillo J M, Kicklighter D W, McGuire A D, Helfrich III J V K, Moore III B, Vörösmarty C J. 1998. Effect of interannual climate variability on carbon storage in Amazonian ecosystems. Nature, 396: 664–667

Tong X, Brandt M, Yue Y M, Ciais P L, Rudbeck Jepsen M, Penuelas J, Wigneron J P, Xiao X M, Song X P, Horion S, Rasmussen K, Saatchi S, Fan L, Wang K L, Zhang B, Chen Z C, Wang Y H, Li X J, Fensholt R. 2020. Forest management in southern China generates short term extensive carbon sequestration. Nat Commun, 11: 129

Trinh D, Trinh Q, Fernández-Cortés A, Mattey D, Guinea J. 2018. First assessment on the air CO 2 dynamic in the show caves of tropical karst, Vietnam. Int J Speleol, 47: 93–112

Walker J C G. 1990. Precambrian evolution of the climate system. Glob Planet Change, 2: 261–289

Wang P, Cao J H, Li L, Yang H, Li G C. 2013. Utilization of Ca 2+ and HCO 3 − in karst water by chlorella from mifferent sources (in Chinese). Acta Hydrobiol Sin, 37: 626–631

Wang S L, Yeager K M, Wan G J, Liu C Q, Wang Y C, Lü Y C. 2012. Carbon export and HCO 3 − fate in carbonate catchments: A case study in the karst plateau of southwestern China. Appl Geochem, 27: 64–72

Wen H, Sullivan P L, Macpherson G L, Billings S A, Li L. 2021. Deepening roots can enhance carbonate weathering by amplifying CO 2 -rich recharge. Biogeosciences, 18: 55–75

White A F, Blum A E. 1995. Effects of climate on chemical weathering in watersheds. Geochim Cosmochim Acta, 59: 1729–1747

Whittaker R H, Likens G E, Woodwell G M, Pecan E V. 1973. Carbon and the Biosphere. In: Proceeding of the 24th Brookhaven Symposium in Biology. Springfield: Technical Information Center, Office of Information Services, US AEC, 1973. 281–302

Xi H P, Wang S J, Bai X Y, Tang H, Luo G J, Li H W, Wu L H, Li C J, Chen H, Ran C, Luo X L. 2021. The responses of weathering carbon sink to eco-hydrological processes in global rocks. Sci Total Environ, 788: 147706

Xiong L, Bai X Y, Zhao C W, Li Y B, Tan Q, Luo G J, Wu L H, Chen F, Li C J, Ran C, Xi H P, Luo X L, Chen H, Zhang S R, Liu M, Gong S H, Xiao B Q, Du C H, Song F J. 2022. High-resolution data sets for global carbonate and silicate rock weathering carbon sinks and their change trends. Earths Future, 10: e2022EF002746

Yang R, Liu Z H, Zeng C, Zhao M. 2012. Response of epikarst hydrochemical changes to soil CO 2 and weather conditions at Chenqi, Puding, SW China. J Hydrol, 468–469: 151–158

Yang Y H, Shi Y, Sun W J, Chang J F, Zhu J X, Chen Y L, Wang X, Guo Y P, Zhang H T, Yu L F, Zhao S Q, Xu H, Zhu J L, Shen H H, Wang Y Y. 2022. Carbon source and sink characteristics of terrestrial ecosystems in China and the world and their contribution to carbon neutrality. Sci China Life Sci, 52: 534–574

Yang M, Liu Z, Sun H, Yang R, Chen B. 2016. Organic carbon source tracing and DIC fertilization effect in the Pearl River: Insights from lipid biomarker and geochemical analysis. Appl GeoChem, 73: 132–141

Yoshimura K, Liu Z H, Cao J H, Yuan D X, Inokura Y, Noto M. 2004. Deep source CO 2 in natural waters and its role in extensive tufa deposition in the Huanglong Ravines, Sichuan, China. Chem Geol, 205: 141–153

Yu G R, Chen Z, Piao S L, Peng C H, Ciais P, Wang Q F, Li X R, Zhu X J. 2014. High carbon dioxide uptake by subtropical forest ecosystems in the East Asian monsoon region. Proc Natl Acad Sci USA, 111: 4910–4915

Yuan D X. 1997. The carbon cycle in karst. Z Geomorphol, 108: 91–102

Zeng C, Wang S, Bai X, Li Y, Tian Y, Li Y, Wu L, Luo G. 2017. Soil erosion evolution and spatial correlation analysis in a typical karst geomorphology using RUSLE with GIS. Solid Earth, 8: 721–736

Zeng S B. 2017. Climate change characteristics of karst area in SW China and its impacts on karst-related carbon sink during recent 40 years (in Chinese). Dissertation for Master’s Degree. Chongqing: Southwest University

Zeng S B, Liu Z H, Kaufmann G. 2019. Sensitivity of the global carbonate weathering carbon-sink flux to climate and land-use changes. Nat Commun, 10: 5749

Zeng S B, Liu Z H, Goldscheider N, Frank S, Goeppert N, Kaufmann G, Zeng C, Zeng Q R, Sun H L. 2021. Comparisons on the effects of temperature, runoff, and landcover on carbonate weathering in different karst catchments: Insights into the future global carbon cycle. Hydrogeol J, 29: 331–345

Zeng S B, Liu Z H. 2022. Karst-related carbon sink and the carbon neutral potential by carbonate liming in non-karst areas in China (in Chinese). Chin Sci Bull, 67: 4116

Zeng S B, Liu Z H, Groves C. 2022. Large-scale CO 2 removal by enhanced carbonate weathering from changes in land-use practices. Earth-Sci Rev, 225: 103915

Zhang C, Pei J G, Xie Y Q, Cao J H, Wang L L. 2008. Impact of land use covers upon karst processes in a typical Fengcong depression system of Nongla, Guangxi, China. Environ Geol, 55: 1621–1626

Zhang C, Xiao Q, Sun P A, Gao X B, Guo Y L, Miao Y, Wang J L. 2022. Progress on karst carbon cycle and carbon sink effect study and perspective (in Chinese). Bull Geol Sci Technol, 41: 190–198

Zhang L K, Qin X Q, Liu P Y, Huang Q B, Lan F, Ji H B. 2014. Estimation of carbon sink fluxes in the Pearl River basin (China) based on a water-rock-gas-organism interaction model. Environ Earth Sci, 74: 945–952

Zhang L K, Qin X Q, Liu P Y, Huang Q B. 2016. Chemical denudation rate and atmospheric CO 2 consumption by H 2 CO 3 and H 2 SO 4 in the Yangtze River Catchment. Acta Geol Sin-Engl, 90: 1933–1944

Zhang S R, Bai X Y, Zhao C W, Tan Q, Luo G J, Wang J F, Li Q, Wu L H, Chen F, Li C J, Deng Y H, Yang Y J, Xi H P. 2021. Global CO 2 consumption by silicate rock chemical weathering: Its past and future. Earths Future, 9: e2020EF001938

Zhang T, Li J H, Pu J B, Yuan D X. 2019. Carbon dioxide exchanges and their controlling factors in Guijiang River, SW China. J Hydrol, 578: 124073

Zomer R J, Bossio D A, Sommer R, Verchot L V. 2017. Global sequestration potential of increased organic carbon in cropland soils. Sci Rep, 7: 15554

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant Nos. U22A20619, 42077455 & 42367008), the Western Light Cross-team Program of Chinese Academy of Sciences (Grant No. xbzg-zdsys-202101), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant Nos. XDB40000000), the Guizhou Provincial Science and Technology Project (Grant Nos. Qiankehe Support [2024] Key 014, Qiankehe Support [2022] Key 010, Qiankehe Support [2023] General 219 & Qiankehe Support ZK(2021)-192), the High-level Innovative Talents in Guizhou Province (Grant No. GCC [2022]015-1), the Opening Fund of the State Key Laboratory of Environmental Geochemistry (Grant No. SKLEG2024202), and the Guizhou Provincial Science and Technology Subsidies (Grant Nos. GZ2019SIG & GZ2020SIG).

Author information

Authors and affiliations.

State Key Laboratory of Environmental Geochemistry, Institute of Geochemistry, Chinese Academy of Sciences, Guiyang, 550081, China

Xiaoyong Bai, Sirui Zhang, Chaojun Li, Lian Xiong, Chaochao Du, Yingying Xue, Zilin Li, Mingkang Long, Minghui Li, Xiaoyun Zhang, Shu Yang, Qing Luo & Xiaoqian Shen

University of Chinese Academy of Sciences, Beijing, 100049, China

Sirui Zhang, Chaojun Li & Mingkang Long

Institute of Biological and Environmental Sciences, University of Aberdeen, Aberdeen, AB24 3UU, UK

College of Resources and Environmental Engineering, Guizhou University, Guiyang, 550025, China

Xiaoyong Bai

School of Geography and Environmental Sciences, Guizhou Normal University, Guiyang, 550025, China

Lian Xiong, Chaochao Du, Yingying Xue, Zilin Li, Minghui Li, Xiaoyun Zhang, Shu Yang, Qing Luo & Xiaoqian Shen

CAS Center for Excellence in Quaternary Science and Global Change, Xi’an, 710061, China

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Xiaoyong Bai .

Ethics declarations

Conflict of interest The authors declare that they have no conflict of interest.

Rights and permissions

Reprints and permissions

About this article

Bai, X., Zhang, S., Smith, P. et al. Resolving controversies surrounding carbon sinks from carbonate weathering. Sci. China Earth Sci. 67 , 2705–2717 (2024). https://doi.org/10.1007/s11430-024-1391-0

Download citation

Received : 19 January 2024

Revised : 17 June 2024

Accepted : 22 July 2024

Published : 08 August 2024

Issue Date : September 2024

DOI : https://doi.org/10.1007/s11430-024-1391-0

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Carbon sink
  • Carbonate weathering
  • Carbon cycle
  • Climate change
  • Anthropogenic driver
  • Find a journal
  • Publish with us
  • Track your research

UTMSI Home

  • Basic Facts
  • Marine Science Advisory Council
  • Recent Highlights
  • Port Aransas Climate
  • Maps & Directions
  • Visit & Learn Home
  • Patton Center
  • Estuary Explorium
  • Wetlands Education Center
  • Bay Education Center
  • Public Programs
  • Field Trips
  • Teacher Workshops
  • Summer Science
  • Scientist in Residence
  • Public Lectures
  • Research Home
  • Fish Physiology and Ecology
  • Ecosystem Dynamics
  • Biogeochemistry
  • Fisheries & Mariculture (FAML)
  • Mission-Aransas NERR
  • Beaufort Lagoon Ecosystems LTER
  • UTMSI Library
  • Technical Seminars
  • Laura Randall Schweppe Endowed Lecture Series in Marine Science
  • Research Vessels
  • Scientific Diving Program
  • Analytical Services Offered
  • Academics Home
  • Undergraduate
  • Course Information
  • Undergraduate Research
  • You are here:  

research about chemical weathering

Weathering Scratches the Surface of Plastic

Weathering Scratches the Surface of Plastic

Plastics might be more resilient than you think—new research shows they can survive in the ocean for decades. A study released in Environmental Science & Technology documents how commercially available macroplastics (>5mm) can withstand decades of exposure in marine environments with minimal degradation. With the influx of plastic waste into the world’s oceans, it is imperative to understand the environmental degradation of plastics and their potential impacts on marine ecosystems.

When plastics are exposed to environmental stressors and mechanical forces, they undergo a process called weathering. This process typically involves the gradual degradation of plastic debris at the surface, leading to increasingly smaller particles until they become virtually indistinguishable. The breakdown of these plastics is known to wreak havoc on marine ecosystems and may also pose risks to human health.

To determine how plastics degrade, researchers at The University of Texas Marine Science Institute along with their colleagues at Ocean Coastal Vision and Marine Biological Laboratory in Woods Hole simulated natural weathering conditions by exposing various types of commercially available macroplastics in seawater to UVA radiation equivalent to 25-75 years of sunlight. Plastics are composed of synthetic polymers, long chains of carbon atoms intertwined with other elements, forming a complex structure. When plastic is exposed to UVA radiation, photodegradation takes place by breaking down the chemical bonds in the polymer. Altering the molecular structure of the polymers can cause cracking, color changes, and loss of physical properties. However, not all plastics respond the same under environmental conditions due to differences in composition.

Various analytical tools were used to characterize the surface chemistry, morphology, thermal stability, and additive composition of plastics. The researchers found that the surface layer of the plastic was highly oxidized and eroded after 9 months of accelerated weathering.  Despite this prolonged exposure and surface layer degradation, the plastics showed little change, retaining both their composition and thermal stability.

Researchers discovered that when the plastics are in the ocean, oxygen-containing compounds will start to form on the surface of plastics. These compounds initiate the degradation process, leading to their breakdown over time. The compounds are formed through reactions with environmental factors like sunlight and oxygen, compromising the structural integrity of the plastics by starting the initial breaking of polymer bonds. However, in some plastics, the presence of antioxidant additives, which are designed to enhance the material’s durability, can inhibit this breakdown. These additives effectively slow down the oxidative degradation process, allowing the plastics to retain their structural integrity for longer periods, even in harsh environmental conditions.

The findings of this study underscore the resilience of macroplastics in marine environments, potentially persisting far longer than previously estimated. This is largely due to the chemical composition of the plastic, the high specific surface area of macroplastics, and the additives the plastic undergoes when created. These factors contribute to the complex interplay between environmental stressors and plastic compositions, influencing the degradation processes. The implications of these findings are significant, suggesting that the long-term impact of plastics on marine ecosystems could be more profound than we currently understand. This research highlights the need for further studies to better understand the fate of plastics in marine ecosystems.

The research was supported by the National Science Foundation and the Simons Foundation. Researchers that conducted the study include Xiangtao Jiang from The University of Texas Marine Science Institute (UTMSI), Scott Gallager from Coastal Ocean Vision, Rut Pedrosa Pàmies from Marine Biological Laboratory (MBL), Emil Ruff from MBL and Zhanfei Liu from UTMSI.

About the author

Amanda R Jacoby

Amanda R Jacoby

Recent posts, related posts, around campus with jace tunnell, utmsi honors advocates, microbes watch out, there’s a new doctor in town, christina bonsell awarded university fellowship, cindy faulk to receive president's outstanding staff award.

College of Natural Sciences The University of Texas at Austin

Web Privacy Policy Web Accessibility Policy

The University of Texas Marine Science Institute 750 Channel View Drive Port Aransas, TX 78373 Telephone Number: 361-749-6711 Fax Number: 361-749-6777

Login » Science and The Sea » Human Resources » MSI Reservations » MSI RTA » Webmail » UT Direct » Sitemap »

UT MSI on Facebook

FAMU seal

Researcher Insights: Environmental Weathering Transforms Plastic Pollution

photo of two plastic bottles floating in water with blue sky

By  showcake  for Adobe Stock

This article is a discussion of scientific research and its potential impacts, prepared by a faculty member at the FAMU-FSU College of Engineering

The prevalence of plastics in our daily lives is reflected by their ubiquity in nature. Jeffrey M. Farner from the FAMU-FSU College of Engineering, details how environmental weathering transforms plastic pollution

Due to use, improper disposal, and environmental transport, plastics are found globally, including in unpopulated locations. This pollution ranges from intact discarded items large enough to be choking hazards for marine life to particles at the nanometer scale. These smallest particles have been observed to translocate into animal tissues, raising concerns about toxicity. The concentrations of plastic particles vary widely, as do the morphology and other characteristics, but they exist.

Environmental stressors lead to transformations

headshot of jeffrey farner phd

Though often discussed as a single material, plastics are a broad classification of products. Further complicating the situation, particles in the environment are exposed to environmental stressors, leading to plastic weathering. Factors like ultraviolet radiation (UV), temperature, abrasion, and shear stresses cause physical and chemical transformations. These result in particles that are generally smaller, more irregularly shaped, and more oxidized than the initial product. (1,2)

Plastic weathering is a function of both the material itself (e.g., polymer, morphology, additives) and the environmental conditions (temperature, UV exposure), and these new particles may behave quite differently compared to the initial engineered plastic. Additionally, the associated changes in hazards arising from weathering still need to be clarified.

Nanoplastic size influences behavior

Particle size influences fate and transport. We have shown that interactions with natural organic matter (NOM) differ for 28 versus 220 nm polystyrene nanoplastics (PSNPs). (3) Using carboxylated polystyrene spheres as models for weathered nanoplastics, we observed aggregation in the presence of NOM and calcium to simulate saline waters. The presence of either alginate or humic acid increased the rate of PSNP aggregation for both particle diameters. Due to bridging, PSNPs were incorporated into NOM-PSNP heteroaggregates, and this effect was more pronounced with suspensions of 220 nm PSNPs.

Curiously, in natural saline water samples, 220 nm PSNPs aggregated more slowly than their smaller counterparts. We hypothesize that the mixture of NOM fractions present may confer stability that is not easily observed in simplified systems. Overall, these results highlight the size-dependency of particle-NOM interactions and the importance of the system itself in which plastic pollution is found.

Surface chemistry plays a role

Changes to the particle surface chemistry can also influence fate and transport. The weathering of polyethylene microspheres due to UV and shear results in fracturing and oxidation. (2) In the context of water treatment, where particle removal is performed via coagulation and flocculation, we observed that increased oxidation of a plastic surface results in greater interactions with coagulant and that removal rates were greater for the smaller, weathered polyethylene particles than for the initial microbeads. These results highlight that the behavior of plastic pollution is impacted by, e.g., size, morphology, and oxidation rather than an intrinsic property solely dependent on the type of material.

Same starting material, different impacts

In addition to influencing fate, weathering-induced transformations can influence the ecotoxicological impact, or hazard, associated with exposure. Recently, we have observed that the same starting material – in this case, polystyrene – will impact sewage sludge in an anaerobic digestor differently based on the path of weathering. (4) Four suspensions of PSNPs were created from the same polystyrene production pellets – spherical, weathered spherical, irregularly shaped, and weathered irregular-shaped particles. Weathering to both spherical and irregularly shaped particles was performed through a combination of heat, UV, and mixing. It resulted in the addition of oxygen-containing functional groups (hydroxyl, carbonyl, and phenolic) at the particle surface. These four suspensions were then introduced at 25 or 150 μg/L to anaerobic digestors to assess how PSNPs interact with microbial communities.

While little change was observed when microbial communities were exposed to 25 μg/L, the addition of 150 μg/L inhibited methane generation, which was attributed to oxidative stress. Irregularly shaped PSNPs (with or without additional weathering) resulted in greater inhibition (up to 20%), which may be due either to their morphology or to an increase in bioavailability. Reactors were sampled for the presence of antibiotic resistance genes (ARGs) and mobile genetic elements, both of which increased in the presence of all PSNPs compared to controls. The abundance of ARGs was further increased with irregularly shaped PSNPs versus spherical and with weathered PSNPs versus unweathered.

These results demonstrate that microbial communities will react differently to the presence of nanoplastics based on particle morphology and oxidation despite originating from the same plastic product. This suggests that the weathering pathway or the history of the plastic itself is important and that effects may be path-dependent.

Weathering complicates our understanding of plastic pollution

Ultimately, this work presents a complex picture in which plastics should not be considered a monolith but rather a classification of pollutants with polymer type as one characteristic. Other similarly important characteristics include particle size and morphology, surface modifications (which can consist of oxidation, surface roughness, sorption, or microbial colonization), and matrix additives. These characteristics will be influenced by the context in which the particles are formed and exist. Differing water chemistries may result in certain characteristics being emphasized over others (e.g., size versus morphology or surface oxidation versus polymer type). Thus, understanding the impact of plastic pollution necessitates identifying the primary drivers of fate, transport, and hazard in a given system.

Hernandez, Laura M., et al. “Analysis of ultraviolet and thermal degradations of four common microplastics and evidence of nanoparticle release.” Journal of Hazardous Materials Letters 4 (2023): 100078.

Lapointe, Mathieu, et al. “Understanding and improving microplastic removal during water treatment: impact of coagulation and flocculation.” Environmental science & technology 54.14 (2020): 8719-8727.

Alimi, Olubukola S., et al. “Mechanistic understanding of the aggregation kinetics of nanoplastics in marine environments: Comparing synthetic and natural water matrices.” Journal of Hazardous Materials Advances 7 (2022): 100115.

Haffiez, Nervana, et al. “Impact of aging of primary and secondary polystyrene nanoplastics on the transmission of antibiotic resistance genes in anaerobic digestion.” Science of The Total Environment (2024): 174213.

This article originally appeared on Open Access Government and is licensed under  Creative Commons Attribution 4.0 International .

  • Share full article

Advertisement

Supported by

For Stonehenge’s Altar Stone, an Improbably Long Ancient Journey

A six-ton megalith at the heart of the archaeological site traveled more than 450 miles to get there, a new study concludes.

A view from some ways away looking at the inner ring of Stonehenge monoliths on an overcast day.

By Franz Lidz

Near the center of the roughly 5,000-year-old circular monument known as Stonehenge is a six-ton, rectangular chunk of red sandstone. In Arthurian legend, the so-called Altar Stone was part of the ring of giant rocks that the wizard Merlin magically transported from Mount Killaurus, in Ireland, to Salisbury Plain, a chalk plateau in southern England — a journey chronicled around 1136 by a Welsh cleric, Geoffrey of Monmouth, in his “ Historia Regum Britanniae .”

Since then, the accepted provenance of the Altar Stone has shifted, spanning a range of possible sites from east Wales and the Marches to northern England. On Wednesday, a study in the journal Nature reroutes the megalith’s odyssey more definitively, proposing a path much longer than scientists had thought possible.

The researchers analyzed the chemical composition and the ages of mineral grains in two microscopic fragments of the Altar Stone. This pinpointed the stone’s source to the Orcadian Basin in northeast Scotland, an area that spans Inverness, the Orkney Islands and Shetland. To reach the archaeological site in Wiltshire, the megalith would have traveled at least 465 miles by land or more than 620 miles along the present-day coastline if it came by sea.

“This is a genuinely shocking result,” said Rob Ixer, a retired mineralogist and research fellow at University College London who collaborated on the project. “The work prompts two important questions: How and why did the stone travel the length of Britain?”

Stonehenge features two kinds of rocks: larger sarsens and smaller bluestones. The sarsens are sandstone slabs found naturally in southern England. They weigh 20 tons on average and were erected in two concentric arrangements. The inner ring is a horseshoe of five trilithons (two uprights capped by a horizontal lintel), of which three complete ones still stand.

We are having trouble retrieving the article content.

Please enable JavaScript in your browser settings.

Thank you for your patience while we verify access. If you are in Reader mode please exit and  log into  your Times account, or  subscribe  for all of The Times.

Thank you for your patience while we verify access.

Already a subscriber?  Log in .

Want all of The Times?  Subscribe .

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 14 August 2024

A Scottish provenance for the Altar Stone of Stonehenge

  • Anthony J. I. Clarke   ORCID: orcid.org/0000-0002-0304-0484 1 ,
  • Christopher L. Kirkland   ORCID: orcid.org/0000-0003-3367-8961 1 ,
  • Richard E. Bevins 2 ,
  • Nick J. G. Pearce   ORCID: orcid.org/0000-0003-3157-9564 2 ,
  • Stijn Glorie 3 &
  • Rob A. Ixer 4  

Nature volume  632 ,  pages 570–575 ( 2024 ) Cite this article

76k Accesses

1 Citations

4486 Altmetric

Metrics details

  • Archaeology

Understanding the provenance of megaliths used in the Neolithic stone circle at Stonehenge, southern England, gives insight into the culture and connectivity of prehistoric Britain. The source of the Altar Stone, the central recumbent sandstone megalith, has remained unknown, with recent work discounting an Anglo-Welsh Basin origin 1 , 2 . Here we present the age and chemistry of detrital zircon, apatite and rutile grains from within fragments of the Altar Stone. The detrital zircon load largely comprises Mesoproterozoic and Archaean sources, whereas rutile and apatite are dominated by a mid-Ordovician source. The ages of these grains indicate derivation from an ultimate Laurentian crystalline source region that was overprinted by Grampian (around 460 million years ago) magmatism. Detrital age comparisons to sedimentary packages throughout Britain and Ireland reveal a remarkable similarity to the Old Red Sandstone of the Orcadian Basin in northeast Scotland. Such a provenance implies that the Altar Stone, a 6 tonne shaped block, was sourced at least 750 km from its current location. The difficulty of long-distance overland transport of such massive cargo from Scotland, navigating topographic barriers, suggests that it was transported by sea. Such routing demonstrates a high level of societal organization with intra-Britain transport during the Neolithic period.

Similar content being viewed by others

research about chemical weathering

The expansion of Acheulean hominins into the Nefud Desert of Arabia

research about chemical weathering

Cryptic geological histories accessed through entombed and matrix geochronometers in dykes

research about chemical weathering

The earliest evidence of Acheulian occupation in Northwest Europe and the rediscovery of the Moulin Quignon site, Somme valley, France

Stonehenge, the Neolithic standing stone circle located on the Salisbury Plain in Wiltshire, England, offers valuable insight into prehistoric Britain. Construction at Stonehenge began as early as 3000  bc , with subsequent modifications during the following two millennia 3 , 4 . The megaliths of Stonehenge are divided into two major categories: sarsen stones and bluestones (Fig. 1a ). The larger sarsens comprise duricrust silcrete predominantly sourced from the West Woods, Marlborough, approximately 25 km north of Stonehenge 5 , 6 . Bluestone, the generic term for rocks considered exotic to the local area, includes volcanic tuff, rhyolite, dolerite and sandstone lithologies 4 (Fig. 1a ). Some lithologies are linked with Neolithic quarrying sites in the Mynydd Preseli area of west Wales 7 , 8 . An unnamed Lower Palaeozoic sandstone, associated with the west Wales area on the basis of acritarch fossils 9 , is present only as widely disseminated debitage at Stonehenge and possibly as buried stumps (Stones 40g and 42c).

figure 1

a , Plan view of Stonehenge showing exposed constituent megaliths and their provenance. The plan of Stonehenge was adapted from ref.  6 under a CC BY 4.0 license. Changes in scale and colour were made, and annotations were added. b , An annotated photograph shows the Altar Stone during a 1958 excavation. The Altar Stone photograph is from the Historic England archive. Reuse is not permitted.

The central megalith of Stonehenge, the Altar Stone (Stone 80), is the largest of the bluestones, measuring 4.9 × 1.0 × 0.5 m, and is a recumbent stone (Fig. 1b ), weighing 6 t and composed of pale green micaceous sandstone with distinctive mineralogy 1 , 2 , 10 (containing baryte, calcite and clay minerals, with a notable absence of K-feldspar) (Fig. 2 ).

figure 2

Minerals with a modal abundance above 0.5% are shown with compositional values averaged across both thin sections. U–Pb ablation pits from laser ablation inductively coupled plasma mass spectrometry (LA-ICP–MS) are shown with age (in millions of years ago, Ma), with uncertainty at the 2 σ level.

Previous petrographic work on the Altar Stone has implied an association to the Old Red Sandstone 10 , 11 , 12 (ORS). The ORS is a late Silurian to Devonian sedimentary rock assemblage that crops out widely throughout Great Britain and Ireland (Extended Data Fig. 1 ). ORS lithologies are dominated by terrestrial siliciclastic sedimentary rocks deposited in continental fluvial, lacustrine and aeolian environments 13 . Each ORS basin reflects local subsidence and sediment infill and thus contains proximal crystalline signatures 13 , 14 .

Constraining the provenance of the Altar Stone could give insights into the connectivity of Neolithic people who left no written record 15 . When the Altar Stone arrived at Stonehenge is uncertain; however, it may have been placed within the central trilithon horseshoe during the second construction phase around 2620–2480  bc 3 . Whether the Altar Stone once stood upright as an approximately 4 m high megalith is unclear 15 ; nevertheless, the current arrangement has Stones 55b and 156 from the collapsed Great Trilithon resting atop the prone and broken Altar Stone (Fig. 1b ).

An early proposed source for the Altar Stone from Mill Bay, Pembrokeshire (Cosheston Subgroup of the Anglo-Welsh ORS Basin), close to the Mynydd Preseli source of the doleritic and rhyolitic bluestones, strongly influenced the notion of a sea transport route via the Bristol Channel 12 . However, inconsistencies in petrography and detrital zircon ages between the Altar Stone and the Cosheston Subgroup have ruled this source out 1 , 11 . Nonetheless, a source from elsewhere in the ORS of the Anglo-Welsh Basin was still considered likely, with an inferred collection and overland transport of the Altar Stone en route to Stonehenge from the Mynydd Preseli 1 . However, a source from the Senni Formation (Cosheston Subgroup) is inconsistent with geochemical and petrographic data, which shows that the Anglo-Welsh Basin is highly unlikely to be the source 2 . Thus, the ultimate provenance of the Altar Stone had remained an open question.

Studies of detrital mineral grains are widely deployed to address questions throughout the Earth sciences and have utility in archaeological investigations 16 , 17 . Sedimentary rocks commonly contain a detrital component derived from a crystalline igneous basement, which may reflect a simple or complex history of erosion, transport and deposition cycles. This detrital cargo can fingerprint a sedimentary rock and its hinterland. More detailed insights become evident when a multi-mineral strategy is implemented, which benefits from the varying degrees of robustness to sedimentary transportation in the different minerals 18 , 19 , 20 .

Here, we present in situ U–Pb, Lu–Hf and trace element isotopic data for zircon, apatite and rutile from two fragments of the Altar Stone collected at Stonehenge: MS3 and 2010K.240 21 , 22 . In addition, we present comparative apatite U–Pb dates for the Orcadian Basin from Caithness and Orkney. We utilize statistical tools (Fig. 3 ) to compare the obtained detrital mineral ages and chemistry (Supplementary Information  1 – 3 ) to crystalline terranes and ORS successions across Great Britain, Ireland and Europe (Fig. 4 and Extended Data Fig. 1 ).

figure 3

a , Multidimensional scaling (MDS) plot of concordant zircon U–Pb ages from the Altar Stone and comparative age datasets, with ellipses at the 95% confidence level 58 . DIM 1 and DIM 2, dimensions 1 and 2. b , Cumulative probability plot of zircon U–Pb ages from crystalline terranes, the Orcadian Basin and the Altar Stone. For a cumulative probability plot of all ORS basins, see Extended Data Fig. 8 .

figure 4

a , Schematic map of Britain, showing outcrops of ORS and other Devonian sedimentary rocks, basement terranes and major faults. Potential Caledonian source plutons are colour-coded on the basis of age 28 . b , Kernel density estimate diagrams displaying zircon U–Pb age (histogram) and apatite Lu–Hf age (dashed line) spectra from the Altar Stone, the Orcadian Basin 25 and plausible crystalline source terranes. The apatite age components for the Altar Stone and Orcadian Basins are shown below their respective kernel density estimates. Extended Data Fig. 3 contains kernel density estimates of other ORS and New Red Sandstone (NRS) age datasets.

Laurentian basement signatures

The crystalline basement terranes of Great Britain and Ireland, from north to south, are Laurentia, Ganderia, Megumia and East Avalonia (Fig. 4a and Extended Data Fig. 1 ). Cadomia-Armorica is south of the Rheic Suture and encompasses basement rocks in western Europe, including northern France and Spain. East Avalonia, Megumia and Ganderia are partly separated by the Menai Strait Fault System (Fig. 4a ). Each terrane has discrete age components, which have imparted palaeogeographic information into overlying sedimentary basins 13 , 14 , 23 . Laurentia was a palaeocontinent that collided with Baltica and Avalonia (a peri-Gondwanan microcontinent) during the early Palaeozoic Caledonian Orogeny to form Laurussia 14 , 24 . West Avalonia is a terrane that includes parts of eastern Canada and comprised the western margin of Avalonia (Extended Data Fig. 1 ).

Statistical comparisons, using a Kolmogorov–Smirnov test, between zircon ages from the Laurentian crystalline basement and the Altar Stone indicate that at a 95% confidence level, no distinction in provenance is evident between Altar Stone detrital zircon U–Pb ages and those from the Laurentian basement. That is, we cannot reject the null hypothesis that both samples are from the same underlying age distribution (Kolmogorov–Smirnov test: P  > 0.05) (Fig. 3a ).

Detrital zircon age components, defined by concordant analyses from at least 4 grains in the Altar Stone, include maxima at 1,047, 1,091, 1,577, 1,663 and 1,790 Ma (Extended Data Fig. 2 ), corresponding to known tectonomagmatic events and sources within Laurentia and Baltica, including the Grenville (1,095–980 Ma), Labrador (1,690–1,590 Ma), Gothian (1,660–1,520 Ma) and Svecokarellian (1,920–1,770 Ma) orogenies 25 .

Laurentian terranes are crystalline lithologies north of the Iapetus Suture Zone (which marks the collision zone between Laurentia and Avalonia) and include the Southern Uplands, Midland Valley, Grampian, Northern Highlands and Hebridean Terranes (Fig. 4a ). Together, these terranes preserve a Proterozoic to Archaean record of zircon production 24 , distinct from the southern Gondwanan-derived terranes of Britain 20 , 26 (Fig. 4a and Extended Data Fig. 3 ).

Age data from Altar Stone rutile grains also point towards an ultimate Laurentian source with several discrete age components (Extended Data Fig. 4 and Supplementary Information  1 ). Group 2 rutile U–Pb analyses from the Altar Stone include Proterozoic ages from 1,724 to 591 Ma, with 3 grains constituting an age peak at 1,607 Ma, overlapping with Laurentian magmatism, including the Labrador and Pinwarian (1,690–1,380 Ma) orogenies 24 . Southern terranes in Britain are not characterized by a large Laurentian (Mesoproterozoic) crystalline age component 25 (Fig. 4b and Extended Data Fig. 3 ). Instead, terranes south of the Iapetus Suture are defined by Neoproterozoic to early Palaeozoic components, with a minor component from around two billion years ago (Figs. 3b and  4b ).

U–Pb analyses of apatite from the Altar Stone define two distinct age groupings. Group 2 apatite U–Pb analyses define a lower intercept age of 1,018 ± 24 Ma ( n  = 9) (Extended Data Fig. 5 ), which overlaps, within uncertainty, to a zircon age component at 1,047 Ma, consistent with a Grenville source 25 . Apatite Lu–Hf dates at 1,496 and 1,151 Ma also imply distinct Laurentian sources 25 (Fig. 4b , Extended Data Fig. 6 and Supplementary Information  2 ). Ultimately, the presence of Grenvillian apatite in the Altar Stone suggests direct derivation from the Laurentian basement, given the lability of apatite during prolonged chemical weathering 20 , 27 .

Grampian Terrane detrital grains

Apatite and rutile U–Pb analyses from the Altar Stone are dominated by regressions from common Pb that yield lower intercepts of 462 ± 4 Ma ( n  = 108) and 451 ± 8 Ma ( n  = 83), respectively (Extended Data Figs. 4 and 5 ). A single concordant zircon analysis also yields an early Palaeozoic age of 498 ± 17 Ma. Hence, with uncertainty from both lower intercepts, Group 1 apatite and rutile analyses demonstrate a mid-Ordovician (443–466 Ma) age component in the Altar Stone. These mid-Ordovician ages are confirmed by in situ apatite Lu–Hf analyses, which define a lower intercept of 470 ± 29 Ma ( n  = 16) (Extended Data Fig. 6 and Supplementary Information  2 ).

Throughout the Altar Stone are sub-planar 100–200-µm bands of concentrated heavy resistive minerals. These resistive minerals are interpreted to be magmatic in origin, given internal textures (oscillatory zonation), lack of mineral overgrowths (in all dated minerals) (Fig. 2 ) and the igneous apatite trace element signatures 27 (Extended Data Fig. 7 and Supplementary Information  3 ). Moreover, there is a general absence of detrital metamorphic zircon grains, further supporting a magmatic origin for these grains.

The most appropriate source region for such mid-Ordovician grains within Laurentian basement is the Grampian Terrane of northeast Scotland (Fig. 4a ). Situated between the Great Glen Fault to the north and the Highland Boundary Fault to the south, the terrane comprises Neoproterozoic to Lower Palaeozoic metasediments termed the Dalradian Supergroup 28 , which are intruded by a compositionally diverse suite of early Palaeozoic granitoids and gabbros (Fig. 4a ). The 466–443 Ma age component from Group 1 apatite and rutile U–Pb analyses overlaps with the terminal stages of Grampian magmatism and subsequent granite pluton emplacement north of the Highland Boundary Fault 28 (Fig. 4a ).

Geochemical classification plots for the Altar Stone apatite imply a compositionally diverse source, much like the lithological diversity within the Grampian Terrane 28 , with 61% of apatite classified as coming from felsic sources, 35% from mafic sources and 4% from alkaline sources (Extended Data Fig. 7 and Supplementary Information  3 ). Specifically, igneous rocks within the Grampian Terrane are largely granitoids, thus accounting for the predominance of felsic-classified apatite grains 29 . We posit that the dominant supply of detritus from 466–443 Ma came from the numerous similarly aged granitoids formed on the Laurentian margin 28 , which are present in both the Northern Highlands and the Grampian Terranes 28 (Fig. 4a ). The alkaline to calc-alkaline suites in these terranes are volumetrically small, consistent with the scarcity of alkaline apatite grains within the Altar Stone (Extended Data Fig. 7 ). Indeed, the Glen Dessary syenite at 447 ± 3 Ma is the only age-appropriate felsic-alkaline pluton in the Northern Highlands Terrane 30 .

The Stacey and Kramers 31 model of terrestrial Pb isotopic evolution predicts a 207 Pb/ 206 Pb isotopic ratio ( 207 Pb/ 206 Pb i ) of 0.8601 for 465 Ma continental crust. Mid-Ordovician regressions through Group 1 apatite and rutile U–Pb analyses yield upper intercepts for 207 Pb/ 206 Pb i of 0.8603 ± 0.0033 and 0.8564 ± 0.0014, respectively (Extended Data Figs. 4 and 5 and Supplementary Information  1 ). The similarity between apatite and rutile 207 Pb/ 206 Pb i implies they were sourced from the same Mid-Ordovician magmatic fluids. Ultimately, the calculated 207 Pb/ 206 Pb i value is consistent with the older (Laurentian) crust north of the Iapetus Suture in Britain 32 (Fig. 4a ).

Orcadian Basin ORS

The detrital zircon age spectra confirm petrographic associations between the Altar Stone and the ORS. Furthermore, the Altar Stone cannot be a New Red Sandstone (NRS) lithology of Permo-Triassic age. The NRS, deposited from around 280–240 Ma, unconformably overlies the ORS 14 . NRS, such as that within the Wessex Basin (Extended Data Fig. 1 ), has characteristic detrital zircon age components, including Carboniferous to Permian zircon grains, which are not present in the Altar Stone 1 , 23 , 26 , 33 , 34 (Extended Data Fig. 3 ).

An ORS classification for the Altar Stone provides the basis for further interpretation of provenance (Extended Data Figs. 1 and 8 ), given that the ORS crops out in distinct areas of Great Britain and Ireland, including the Anglo-Welsh border and south Wales, the Midland Valley and northeast Scotland, reflecting former Palaeozoic depocentres 14 (Fig. 4a ).

Previously reported detrital zircon ages and petrography show that ORS outcrops of the Anglo-Welsh Basin in the Cosheston Subgroup 1 and Senni Formation 2 are unlikely to be the sources of the Altar Stone (Fig. 4a ). ORS within the Anglo-Welsh Basin is characterized by mid-Palaeozoic zircon age maxima and minor Proterozoic components (Fig. 4a ). Ultimately, the detrital zircon age spectra of the Altar Stone are statistically distinct from the Anglo-Welsh Basin (Fig. 3a ). In addition, the ORS outcrops of southwest England (that is, south of the Variscan front), including north Devon and Cornwall (Cornubian Basin) (Fig. 4a ), show characteristic facies, including marine sedimentary structures and fossils along with a metamorphic fabric 13 , 26 , inconsistent with the unmetamorphosed, terrestrial facies of the Altar Stone 1 , 11 .

Another ORS succession with published age data for comparison is the Dingle Peninsula Basin, southwest Ireland. However, the presence of late Silurian (430–420 Ma) and Devonian (400–350 Ma) apatite, zircon and muscovite from the Dingle Peninsula ORS discount a source for the Altar Stone from southern Ireland 20 . The conspicuous absence of apatite grains of less than 450 Ma in age in the Altar Stone precludes the input of Late Caledonian magmatic grains to the source sediment of the Altar Stone and demonstrates that the ORS of the Altar Stone was deposited prior to or distally from areas of Late Caledonian magmatism, unlike the ORS of the Dingle Peninsula 20 . Notably, no distinction in provenance between the Anglo-Welsh Basin and the Dingle Peninsula ORS is evident (Kolmogorov–Smirnov test: P  > 0.05), suggesting that ORS basins south of the Iapetus Suture are relatively more homogenous in terms of their detrital zircon age components (Fig. 4a ).

In Scotland, ORS predominantly crops out in the Midland Valley and Orcadian Basins (Fig. 4a ). The Midland Valley Basin is bound between the Highland Boundary Fault and the Iapetus Suture and is located within the Midland Valley and Southern Uplands Terranes. Throughout Midland Valley ORS stratigraphy, detrital zircon age spectra broadly show a bimodal age distribution between Lower Palaeozoic and Mesoproterozoic components 35 , 36 (Extended Data Fig. 3 ). Indeed, throughout 9 km of ORS stratigraphy in the Midland Valley Basin and across the Sothern Uplands Fault, no major changes in provenance are recognized 36 (Fig. 4a ). Devonian zircon, including grains as young as 402 ± 5 Ma from the northern ORS in the Midland Valley Basin 36 , further differentiates this basin from the Altar Stone (Fig. 3a and Extended Data Fig. 3 ). The scarcity of Archaean to late Palaeoproterozoic zircon grains within the Midland Valley ORS shows that the Laurentian basement was not a dominant detrital source for those rocks 35 . Instead, ORS of the Midland Valley is primarily defined by zircon from 475 Ma interpreted to represent the detrital remnants of Ordovician volcanism within the Midland Valley Terrane, with only minor and periodic input from Caledonian plutonism 35 .

The Orcadian Basin of northeast Scotland, within the Grampian and Northern Highlands terranes, contains a thick package of mostly Mid-Devonian ORS, around 4 km thick in Caithness and up to around 8 km thick in Shetland 14 (Fig. 4a ). The detrital zircon age spectra from Orcadian Basin ORS provides the closest match to the Altar Stone detrital ages 25 (Fig. 3 and Extended Data Fig. 8 ). A Kolmogorov–Smirnov test on age spectra from the Altar Stone and the Orcadian Basin fails to reject the null hypothesis that they are derived from the same underlying distribution (Kolmogorov–Smirnov test: P  > 0.05) (Fig. 3a ). To the north, ORS on the Svalbard archipelago formed on Laurentian and Baltican basement rocks 37 . Similar Kolmogorov–Smirnov test results, where each detrital zircon dataset is statistically indistinguishable, are obtained for ORS from Svalbard, the Orcadian Basin and the Altar Stone.

Apatite U–Pb age components from Orcadian Basin samples from Spittal, Caithness (AQ1) and Cruaday, Orkney (CQ1) (Fig. 4a ) match those from the Altar Stone. Group 2 apatite from the Altar Stone at 1,018 ± 24 Ma is coeval with a Grenvillian age from Spittal at 1,013 ± 35 Ma. Early Palaeozoic apatite components at 473 ± 25 Ma and 466 ± 6 Ma, from Caithness and Orkney, respectively (Extended Data Fig. 5 and Supplementary Information  1 ), are also identical, within uncertainty, to Altar Stone Group 1 (462 ± 4 Ma) apatite U–Pb analyses and a Lu–Hf component at 470 ± 28 Ma supporting a provenance from the Orcadian Basin for the Altar Stone (Extended Data Fig. 6 and Supplementary Information  2 ).

During the Palaeozoic, the Orcadian Basin was situated between Laurentia and Baltica on the Laurussian palaeocontinent 14 . Correlations between detrital zircon age components imply that both Laurentia and Baltica supplied sediment into the Orcadian Basin 25 , 36 . Detrital grains from more than 900 Ma within the Altar Stone are consistent with sediment recycling from intermediary Neoproterozoic supracrustal successions (for example, Dalradian Supergroup) within the Grampian Terrane but also from the Särv and Sparagmite successions of Baltica 25 , 36 . At around 470 Ma, the Grampian Terrane began to denude 28 . Subsequently, first-cycle detritus, such as that represented by Group 1 apatite and rutile, was shed towards the Orcadian Basin from the southeast 25 .

Thus, the resistive mineral cargo in the Altar Stone represents a complex mix of first and multi-cycle grains from multiple sources. Regardless of total input from Baltica versus Laurentia into the Orcadian Basin, crystalline terranes north of the Iapetus Suture (Fig. 4a ) have distinct age components that match the Altar Stone in contrast to Gondwanan-derived terranes to the south.

The Altar Stone and Neolithic Britain

Isotopic data for detrital zircon and rutile (U–Pb) and apatite (U–Pb, Lu–Hf and trace elements) indicate that the Altar Stone of Stonehenge has a provenance from the ORS in the Orcadian Basin of northeast Scotland (Fig. 4a ). Given this detrital mineral provenance, the Altar Stone cannot have been sourced from southern Britain (that is, south of the Iapetus Suture) (Fig. 4a ), including the Anglo-Welsh Basin 1 , 2 .

Some postulate a glacial transport mechanism for the Mynydd Preseli (Fig. 4a ) bluestones to Salisbury Plain 38 , 39 . However, such transport for the Altar Stone is difficult to reconcile with ice-sheet reconstructions that show a northwards movement of glaciers (and erratics) from the Grampian Mountains towards the Orcadian Basin during the Last Glacial Maximum and, indeed, previous Pleistocene glaciations 40 , 41 . Moreover, there is little evidence of extensive glacial deposition in central southern Britain 40 , nor are Scottish glacial erratics found at Stonehenge 42 . Sr and Pb isotopic signatures from animal and human remains from henges on Salisbury Plain demonstrate the mobility of Neolithic people within Britain 32 , 43 , 44 , 45 . Furthermore, shared architectural elements and rock art motifs between Neolithic monuments in Orkney, northern Britain, and Ireland point towards the long-distance movement of people and construction materials 46 , 47 .

Thus, we posit that the Altar Stone was anthropogenically transported to Stonehenge from northeast Scotland, consistent with evidence of Neolithic inhabitation in this region 48 , 49 . Whereas the igneous bluestones were brought around 225 km from the Mynydd Preseli to Stonehenge 50 (Fig. 4a ), a Scottish provenance for the Altar Stone demands a transport distance of at least 750 km (Fig. 4a ). Nonetheless, even with assistance from beasts of burden 51 , rivers and topographical barriers, including the Grampians, Southern Uplands and the Pennines, along with the heavily forested landscape of prehistoric Britain 52 , would have posed formidable obstacles for overland megalith transportation.

At around 5000  bc , Neolithic people introduced the common vole ( Microtus arvalis ) from continental Europe to Orkney, consistent with the long-distance marine transport of cattle and goods 53 . A Neolithic marine trade network of quarried stone tools is found throughout Britain, Ireland and continental Europe 54 . For example, a saddle quern, a large stone grinding tool, was discovered in Dorset and determined to have a provenance in central Normandy 55 , implying the shipping of stone cargo over open water during the Neolithic. Furthermore, the river transport of shaped sandstone blocks in Britain is known from at least around 1500  bc (Hanson Log Boat) 56 . In Britain and Ireland, sea levels approached present-day heights from around 4000  bc 57 , and although coastlines have shifted, the geography of Britain and Ireland would have permitted sea routes southward from the Orcadian Basin towards southern England (Fig. 4a ). A Scottish provenance for the Altar Stone implies Neolithic transport spanning the length of Great Britain.

This work analysed two 30-µm polished thin sections of the Altar Stone (MS3 and 2010K.240) and two sections of ORS from northeast Scotland (Supplementary Information  4 ). CQ1 is from Cruaday, Orkney (59° 04′ 34.2″ N, 3° 18′ 54.6″ W), and AQ1 is from near Spittal, Caithness (58° 28′ 13.8″ N, 3° 27′ 33.6″ W). Conventional optical microscopy (transmitted and reflected light) and automated mineralogy via a TESCAN Integrated Mineral Analyser gave insights into texture and mineralogy and guided spot placement during LA-ICP–MS analysis. A CLARA field emission scanning electron microscope was used for textural characterization of individual minerals (zircon, apatite and rutile) through high-resolution micrometre-scale imaging under both back-scatter electron and cathodoluminescence. The Altar Stone is a fine-grained and well-sorted sandstone with a mean grain size diameter of ≤300 µm. Quartz grains are sub-rounded and monocrystalline. Feldspars are variably altered to fine-grained white mica. MS3 and 2010K.240 have a weakly developed planar fabric and non-planar heavy mineral laminae approximately 100–200 µm thick. Resistive heavy mineral bands are dominated by zircon, rutile, and apatite, with grains typically 10–40 µm wide. The rock is mainly cemented by carbonate, with localized areas of barite and quartz cement. A detailed account of Altar Stone petrography is provided in refs. 1 , 59 .

Zircon isotopic analysis

Zircon u–pb methods.

Two zircon U–Pb analysis sessions were completed at the GeoHistory facility in the John De Laeter Centre (JdLC), Curtin University, Australia. Ablations within zircon grains were created using an excimer laser RESOlution LE193 nm ArF with a Laurin Technic S155 cell. Isotopic data was collected with an Agilent 8900 triple quadrupole mass spectrometer, with high-purity Ar as the plasma carrier gas (flow rate 1.l min −1 ). An on-sample energy of ~2.3–2.7 J cm −2 with a 5–7 Hz repetition rate was used to ablate minerals for 30–40 s (with 25–60 s of background capture). Two cleaning pulses preceded analyses, and ultra-high-purity He (0.68 ml min −1 ) and N 2 (2.8 ml min −1 ) were used to flush the sample cell. A block of reference mineral was analysed following 15–20 unknowns. The small, highly rounded target grains of the Altar Stone (usually <30 µm in width) necessitated using a spot size diameter of ~24 µm for all ablations. Isotopic data was reduced using Iolite 4 60 with the U-Pb Geochronology data reduction scheme, followed by additional calculation and plotting via IsoplotR 61 . The primary matrix-matched reference zircon 62 used to correct instrumental drift and mass fractionation was GJ-1, 601.95 ± 0.40 Ma. Secondary reference zircon included Plešovice 63 , 337.13 ± 0.37 Ma, 91500 64 , 1,063.78 ± 0.65 Ma, OG1 65 3,465.4 ± 0.6 Ma and Maniitsoq 66 3,008.7 ± 0.6 Ma. Weighted mean U–Pb ages for secondary reference materials were within 2 σ uncertainty of reported values (Supplementary Information  5 ).

Zircon U–Pb results

Across two LA-ICP–MS sessions, 83 U–Pb measurements were obtained on as many zircon grains; 41 were concordant (≤10% discordant), where discordance is defined using the concordia log distance (%) approach 67 . We report single-spot (grain) concordia ages, which have numerous benefits over conventional U–Pb/Pb–Pb ages, including providing an objective measure of discordance that is directly coupled to age and avoids the arbitrary switch between 206 Pb/ 238 U and 207 Pb/ 206 Pb. Furthermore, given the spread in ages (Early Palaeozoic to Archaean), concordia ages provide optimum use of both U–Pb/Pb–Pb ratios, offering greater precision over 206 Pb/ 238 U or 207 Pb/ 206 Pb ages alone.

Given that no direct sampling of the Altar Stone is permitted, we are limited in the amount of material available for destructive analysis, such as LA-ICP–MS. We collate our zircon age data with the U–Pb analyses 1 of FN593 (another fragment of the Altar Stone), filtered using the same concordia log distance (%) discordance filter 67 . The total concordant analyses used in this work is thus 56 over 3 thin sections, each showing no discernible provenance differences. Zircon concordia ages span from 498 to 2,812 Ma. Age maxima (peak) were calculated after Gehrels 68 , and peak ages defined by ≥4 grains include 1,047, 1,091, 1,577, 1,663 and 1,790 Ma.

For 56 concordant ages from 56 grains at >95% certainty, the largest unmissed fraction is calculated at 9% of the entire uniform detrital population 69 . In any case, the most prevalent and hence provenance important components will be sampled for any number of analyses 69 . We analysed all zircon grains within the spatial limit of the technique in the thin sections 70 . We used in situ thin-section analysis, which can mitigate against contamination and sampling biases in detrital studies 71 . Adding apatite (U–Pb and Lu–Hf) and rutile (U–Pb) analyses bolsters our confidence in provenance interpretations as these minerals will respond dissimilarly during transport.

Comparative zircon datasets

Zircon U–Pb compilations of the basement terranes of Britain and Ireland were sourced from refs. 20 , 26 . ORS detrital zircon datasets used for comparison include isotopic data from the Dingle Peninsula Basin 20 , Anglo-Welsh Basin 72 , Midland Valley Basin 35 , Svalbard ORS 37 and Orcadian Basin 25 . NRS zircon U–Pb ages were sourced from the Wessex Basin 33 . Comparative datasets were filtered for discordance as per our definition above 20 , 26 . Kernel density estimates for age populations were created within IsoplotR 61 using a kernel and histogram bandwidth of 50 Ma.

A two-sample Kolmogorov–Smirnov statistical test was implemented to compare the compiled zircon age datasets with the Altar Stone (Supplementary Information  6 ). This two-sided test compares the maximum probability difference between two cumulative density age functions, evaluating the null hypothesis that both age spectra are drawn from the same distribution based on a critical value dependent on the number of analyses and a chosen confidence level.

The number of zircon ages within the comparative datasets used varies from the Altar Stone ( n  = 56) to Laurentia ( n  = 2,469). Therefore, to address the degree of dependence on sample n , we also implemented a Monte Carlo resampling (1,000 times) procedure for the Kolmogorov–Smirnov test, including the uncertainty on each age determination to recalculate P values and standard deviations (Supplementary Information  7 ), based on the resampled distribution of each sample. The results from Kolmogorov–Smirnov tests, using Monte Carlo resampling (and multidimensional analysis), taking uncertainty due to sample n into account, also support the interpretation that at >95% certainty, no distinction in provenance can be made between the Altar Stone zircon age dataset ( n  = 56) and those from the Orcadian Basin ( n  = 212), Svalbard ORS ( n  = 619 ) and the Laurentian basement (Supplementary Information  7 ).

MDS plots for zircon datasets were created using the MATLAB script of ref.  58 . Here, we adopted a bootstrap resampling (>1,000 times) with Procrustes rotation of Kolmogorov–Smirnov values, which outputs uncertainty ellipses at a 95% confidence level (Fig. 3a ). In MDS plots, stress is a goodness of fit indicator between dissimilarities in the datasets and distances on the MDS plot. Stress values below 0.15 are desirable 58 . For the MDS plot in Fig. 3a , the value is 0.043, which indicates an “excellent” fit 58 .

Rutile isotopic analysis

Rutile u–pb methods.

One rutile U–Pb analysis session was completed at the GeoHistory facility in the JdLC, Curtin University, Australia. Rutile grains were ablated (24 µm) using a Resonetics RESOlution M-50A-LR sampling system, using a Compex 102 excimer laser, and measured using an Agilent 8900 triple quadrupole mass analyser. The analytical parameters included an on-sample energy of 2.7 J cm −2 , a repetition rate of 7 Hz for a total analysis time of 45 s, and 60 s of background data capture. The sample chamber was purged with ultrahigh purity He at a flow rate of 0.68 l min −1 and N 2 at 2.8 ml min −1 .

U–Pb data for rutile analyses was reduced against the R-10 rutile primary reference material 73 (1,091 ± 4 Ma). The secondary reference material used to monitor the accuracy of U–Pb ratios was R-19 rutile. The mean weighted 238 U/ 206 Pb age obtained for R-19 was 491 ± 10 (mean squared weighted deviation (MSWD) = 0.87, p ( χ 2 ) = 0.57) within uncertainty of the accepted age 74 of 489.5 ± 0.9 Ma.

Rutile grains with negligible Th concentrations can be corrected for common Pb using a 208 Pb correction 74 . Previously used thresholds for Th content have included 75 , 76 Th/U < 0.1 or a Th concentration >5% U. However, Th/U ratios for rutile from MS3 are typically > 1; thus, a 208 Pb correction is not applicable. Instead, we use a 207 -based common Pb correction 31 to account for the presence of common Pb. Rutile isotopic data was reduced within Iolite 4 60 using the U–Pb Geochronology reduction scheme and IsoplotR 61 .

Rutile U–Pb Results

Ninety-two rutile U–Pb analyses were obtained in a U–Pb single session, which defined two coherent age groupings on a Tera–Wasserburg plot.

Group 1 constitutes 83 U–Pb rutile analyses, forming a well-defined mixing array on a Tera-Wasserburg plot between common and radiogenic Pb components. This array yields an upper intercept of 207 Pb/ 206 Pb i  = 0.8563 ± 0.0014. The lower intercept implies an age of 451 ± 8 Ma. The scatter about the line (MSWD = 2.7) is interpreted to reflect the variable passage of rutile of diverse grain sizes through the radiogenic Pb closure temperature at ~600 °C during and after magmatic crystallization 77 .

Group 2 comprises 9 grains, with 207 Pb corrected 238 U/ 206 Pb ages ranging from 591–1,724 Ma. Three grains from Group 2 define an age peak 68 at 1,607 Ma. Given the spread in U–Pb ages, we interpret these Proterozoic grains to represent detrital rutile derived from various sources.

Apatite isotopic analysis

Apatite u–pb methods.

Two apatite U–Pb LA-ICP–MS analysis sessions were conducted at the GeoHistory facility in the JdLC, Curtin University, Australia. For both sessions, ablations were created using a RESOlution 193 nm excimer laser ablation system connected to an Agilent 8900 ICP–MS with a RESOlution LE193 nm ArF and a Laurin Technic S155 cell ICP–MS. Other analytical details include a fluence of 2 J cm 2 and a 5 Hz repetition rate. For the Altar Stone section (MS3) and the Orcadian Basin samples (Supplementary Information  4 ), 24- and 20-µm spot sizes were used, respectively.

The matrix-matched primary reference material used for apatite U–Pb analyses was the Madagascar apatite (MAD-1) 78 . A range of secondary reference apatite was analysed, including FC-1 79 (Duluth Complex) with an age of 1,099.1 ± 0.6 Ma, Mount McClure 80 , 81 526 ± 2.1 Ma, Otter Lake 82 913 ± 7 Ma and Durango 31.44 ± 0.18 83  Ma. Anchored regressions (through reported 207 Pb/ 206 Pb i values) for secondary reference material yielded lower intercept ages within 2 σ uncertainty of reported values (Supplementary Information  8 ).

Altar Stone apatite U–Pb results

This first session of apatite U–Pb of MS3 from the Altar Stone yielded 117 analyses. On a Tera–Wasserburg plot, these analyses form two discordant mixing arrays between common and radiogenic Pb components with distinct lower intercepts.

The array from Group 2 apatite, comprised of 9 analyses, yields a lower intercept equivalent to an age of 1,018 ± 24 Ma (MSWD = 1.4) with an upper intercept 207 Pb/ 206 Pb i  = 0.8910 ± 0.0251. The f 207 % (the percentage of common Pb estimated using the 207 Pb method) of apatite analyses in Group 2 ranges from 16.66–88.8%, with a mean of 55.76%.

Group 1 apatite is defined by 108 analyses yielding a lower intercept of 462 ± 4 Ma (MSWD = 2.4) with an upper intercept 207 Pb/ 206 Pb i  = 0.8603 ± 0.0033. The f 207 % of apatite analyses in Group 1 range from 10.14–99.91%, with a mean of 78.65%. The slight over-dispersion of the apatite regression line may reflect some variation in Pb closure temperature in these crystals 84 .

Orcadian basin apatite U–Pb results

The second apatite U–Pb session yielded 138 analyses from samples CQ1 and AQ1. These data form three discordant mixing arrays between radiogenic and common Pb components on a Tera–Wasserburg plot.

An unanchored regression through Group 1 apatite ( n  = 14) from the Cruaday sample (CQ1) yields a lower intercept of 473 ± 25 Ma (MSWD = 1.8) with an upper intercept of 207 Pb/ 206 Pb i  = 0.8497 ± 0.0128. The f 207 % spans 38–99%, with a mean value of 85%.

Group 1 from the Spittal sample (AQ1), comprised of 109 analyses, yields a lower intercept equal to 466 ± 6 Ma (MSWD = 1.2). The upper 207 Pb/ 206 Pb i is equal to 0.8745 ± 0.0038. f 207 % values for this group range from 6–99%, with a mean value of 83%. A regression through Group 2 analyses ( n  = 17) from the Spittal sample yields a lower intercept of 1,013 ± 35 Ma (MSWD = 1) and an upper intercept 207 Pb/ 206 Pb i of 0.9038 ± 0.0101. f 207 % values span 25–99%, with a mean of 76%. Combined U–Pb analyses from Groups 1 from CQ1 and AQ1 ( n  = 123) yield a lower intercept equivalent to 466 ± 6 Ma (MSWD = 1.4) and an upper intercept 207 Pb/ 206 Pb i of 0.8726 ± 0.0036, which is presented beneath the Orcadian Basin kernel density estimate in Fig. 4b .

Apatite Lu–Hf methods

Apatite grains were dated in thin-section by the in situ Lu–Hf method at the University of Adelaide, using a RESOlution-LR 193 nm excimer laser ablation system, coupled to an Agilent 8900 ICP–MS/MS 85 , 86 . A gas mixture of NH 3 in He was used in the mass spectrometer reaction cell to promote high-order Hf reaction products, while equivalent Lu and Yb reaction products were negligible. The mass-shifted (+82 amu) reaction products of 176+82 Hf and 178+82 Hf reached the highest sensitivity of the measurable range and were analysed free from isobaric interferences. 177 Hf was calculated from 178 Hf, assuming natural abundances. 175 Lu was measured on mass as a proxy 85 for 176 Lu. Laser ablation was conducted with a laser beam of 43 µm at 7.5 Hz repetition rate and a fluency of approximately 3.5 J cm −2 . The analysed isotopes (with dwell times in ms between brackets) are 27 Al (2), 43 Ca (2), 57 Fe (2), 88 Sr (2), 89+85 Y (2), 90+83 Zr (2), 140+15 Ce (2), 146 Nd (2), 147 Sm (2), 172 Yb (5), 175 Lu (10), 175+82 Lu (50), 176+82 Hf (200) and 178+82 Hf (150). Isotopes with short dwell times (<10 ms) were measured to confirm apatite chemistry and to monitor for inclusions. 175+82 Lu was monitored for interferences on 176+82 Hf.

Relevant isotope ratios were calculated in LADR 87 using NIST 610 as the primary reference material 88 . Subsequently, reference apatite OD-306 78 (1,597 ± 7 Ma) was used to correct the Lu–Hf isotope ratios for matrix-induced fractionation 86 , 89 . Reference apatites Bamble-1 (1,597 ± 5 Ma), HR-1 (344 ± 2 Ma) and Wallaroo (1,574 ± 6 Ma) were monitored for accuracy verification 85 , 86 , 90 . Measured Lu–Hf dates of 1,098 ± 7 Ma, 346.0 ± 3.7 Ma and 1,575 ± 12 Ma, respectively, are in agreement with published values. All reference materials have negligible initial Hf, and weighted mean Lu–Hf dates were calculated in IsoplotR 61 directly from the (matrix-corrected) 176 Hf/ 176 Lu ratios.

For the Altar Stone apatites, which have variable 177 Hf/ 176 Hf compositions, single-grain Lu–Hf dates were calculated by anchoring isochrons to an initial 177 Hf/ 176 Hf composition 90 of 3.55 ± 0.05, which spans the entire range of initial 177 Hf/ 176 Hf ratios of the terrestrial reservoir (for example, ref. 91 ). The reported uncertainties for the single-grain Lu–Hf dates are presented as 95% confidence intervals, and dates are displayed on a kernel density estimate plot.

Apatite Lu–Hf results

Forty-five apatite Lu–Hf analyses were obtained from 2010K.240. Those with radiogenic Lu ingrowth or lacking common Hf gave Lu–Hf ages, defining four coherent isochrons and age groups.

Group 1, defined by 16 grains, yields a Lu–Hf isochron with a lower intercept of 470 ± 28 Ma (MSWD = 0.16, p ( χ 2 ) = 1). A second isochron through 5 analyses (Group 2) constitutes a lower intercept equivalent to 604 ± 38 Ma (MSWD = 0.14, p ( χ 2 ) = 0.94). Twelve apatite Lu–Hf analyses define Group 3 with a lower intercept of 1,123 ± 42 Ma (MSWD = 0.75, p ( χ 2 ) = 0.68). Three grains constitute the oldest grouping, Group 4 at 1,526 ± 186 Ma (MSWD = 0.014, p ( χ 2 ) = 0.91).

Apatite trace elements methods

A separate session of apatite trace element analysis was undertaken. Instrumentation and analytical set-up were identical to that described in 4.1. NIST 610 glass was the primary reference material for apatite trace element analyses. 43 Ca was used as the internal reference isotope, assuming an apatite Ca concentration of 40 wt%. Secondary reference materials included NIST 612 and the BHVO−2g glasses 92 . Elemental abundances for secondary reference material were generally within 5–10% of accepted values. Apatite trace element data was examined using the Geochemical Data Toolkit 93 .

Apatite trace elements results

One hundred and thirty-six apatite trace element analyses were obtained from as many grains. Geochemical classification schemes for apatite were used 29 , and three compositional groupings (felsic, mafic-intermediate, and alkaline) were defined.

Felsic-classified apatite grains ( n  = 83 (61% of analyses)) are defined by La/Nd of <0.6 and (La + Ce + Pr)/ΣREE (rare earth elements) of <0.5. The median values of felsic grains show a flat to slightly negative gradient on the chondrite-normalized REE plot from light to heavy REEs 94 . Felsic apatite’s median europium anomaly (Eu/Eu*) is 0.59, a moderately negative signature.

Mafic-intermediate apatite 29 ( n  = 48 (35% of grains)) are defined by (La + Ce + Pr)/ΣREE of 0.5–0.7 and a La/Nd of 0.5–1.5. In addition, apatite grains of this group typically exhibit a chondrite-normalized Ce/Yb of >5 and ΣREEs up to 1.25 wt%. Apatite grains classified as mafic-intermediate show a negative gradient on a chondrite-normalized REE plot from light to heavy REEs. The apatite grains of this group generally show the most enrichment in REEs compared to chondrite 94 . The median europium (Eu/Eu*) of mafic-intermediate apatite is 0.62, a moderately negative anomaly.

Lastly, alkaline apatite grains 29 ( n  = 5 (4% of analyses)) are characterized by La/Nd > 1.5 and a (La + Ce + Pr)/ΣREE > 0.8. The median europium anomaly of this group is 0.45. This grouping also shows elevated chondrite-normalized Ce/Yb of >10 and >0.5 wt% for the ΣREEs.

Reporting summary

Further information on research design is available in the  Nature Portfolio Reporting Summary linked to this article.

Data availability

The isotopic and chemical data supporting the findings of this study are available within the paper and its supplementary information files.

Bevins, R. E. et al. Constraining the provenance of the Stonehenge ‘Altar Stone’: evidence from automated mineralogy and U–Pb zircon age dating. J. Archaeolog. Sci. 120 , 105188 (2020).

Article   CAS   Google Scholar  

Bevins, R. E. et al. The Stonehenge Altar Stone was probably not sourced from the Old Red Sandstone of the Anglo-Welsh Basin: time to broaden our geographic and stratigraphic horizons? J. Archaeolog. Sci. Rep. 51 , 104215 (2023).

Google Scholar  

Pearson, M. P. et al. in Stonehenge for the Ancestors: Part 2: Synthesis (eds Pearson, M. P. et al.) 47–75 (Sidestone Press, 2022).

Pitts, M. W. How to Build Stonehenge (Thames & Hudson, 2022).

Nash, D. J. et al. Origins of the sarsen megaliths at Stonehenge. Sci. Adv. 6 , eabc0133 (2020).

Article   ADS   CAS   PubMed   PubMed Central   Google Scholar  

Nash, D. J. et al. Petrological and geochemical characterisation of the sarsen stones at Stonehenge. PLoS ONE 16 , e0254760 (2021).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Pearson, M. P. et al. Megalith quarries for Stonehenge’s bluestones. Antiquity 93 , 45–62 (2019).

Article   Google Scholar  

Pearson, M. P. et al. Craig Rhos-y-felin: a Welsh bluestone megalith quarry for Stonehenge. Antiquity 89 , 1331–1352 (2015).

Ixer, R., Turner, P., Molyneux, S. & Bevins, R. The petrography, geological age and distribution of the Lower Palaeozoic Sandstone debitage from the Stonehenge landscape. Wilts. Archaeol. Nat. Hist. Mag. 110 , 1–16 (2017).

Ixer, R. & Turner, P. A detailed re-examination of the petrography of the Altar Stone and other non-sarsen sandstones from Stonehenge as a guide to their provenance. Wilts. Archaeol. Nat. Hist. Mag. 99 , 1–9 (2006).

Ixer, R., Bevins, R. E., Pirrie, D., Turner, P. & Power, M. No provenance is better than wrong provenance: Milford Haven and the Stonehenge sandstones. Wilts. Archaeol. Nat. Hist. Mag. 113 , 1–15 (2020).

Thomas, H. H. The source of the stones of Stonehenge. The Antiq. J. 3 , 239–260 (1923).

Kendall, R. S. The Old Red Sandstone of Britain and Ireland—a review. Proc. Geol. Assoc. 128 , 409–421 (2017).

Woodcock, N., Holdsworth, R. E. & Strachan, R. A. in Geological History of Britain and Ireland (eds Woodcock, N. & Strachan, R. A.) Ch. 6 91–109 (Wiley-Blackwell, 2012).

Pearson, M. P., Pollard, J., Richards, C., Thomas, J. & Welham, K. Stonehenge: Making Sense of a Prehistoric Mystery (Council for British Archaeology, 2015).

Shewan, L. et al. Dating the megalithic culture of laos: Radiocarbon, optically stimulated luminescence and U/Pb zircon results. PLoS ONE 16 , e0247167 (2021).

Kelloway, S. et al. Sourcing olive jars using U–Pb ages of detrital zircons: a study of 16th century olive jars recovered from the Solomon Islands. Geoarchaeology 29 , 47–60 (2014).

Barham, M. et al. The answers are blowin’ in the wind: ultra-distal ashfall zircons, indicators of Cretaceous super-eruptions in eastern Gondwana. Geology 44 , 643–646 (2016).

Article   ADS   CAS   Google Scholar  

Gillespie, J., Glorie, S., Khudoley, A. & Collins, A. S. Detrital apatite U–Pb and trace element analysis as a provenance tool: Insights from the Yenisey Ridge (Siberia). Lithos 314–315 , 140–155 (2018).

Article   ADS   Google Scholar  

Fairey, B. J. et al. The provenance of the Devonian Old Red Sandstone of the Dingle Peninsula, SW Ireland; the earliest record of Laurentian and peri-Gondwanan sediment mixing in Ireland. J. Geol. Soc. 175 , 411–424 (2018).

Bevins, R. E. et al. Assessing the authenticity of a sample taken from the Altar Stone at Stonehenge in 1844 using portable XRF and automated SEM-EDS. J. Archaeol. Sci. Rep. 49 , 103973 (2023).

Bevins, R. E. et al. Linking derived debitage to the Stonehenge Altar Stone using portable X-ray fluorescence analysis. Mineral. Mag. 86 , 688–700 (2022).

Morton, A. C., Chisholm, J. I. & Frei, D. Provenance of Carboniferous sandstones in the central and southern parts of the Pennine Basin, UK: evidence from detrital zircon ages. Proc. York. Geol. Soc. 63 , https://doi.org/10.1144/pygs2020-010 (2021).

Cawood, P. A., Nemchin, A. A., Strachan, R., Prave, T. & Krabbendam, M. Sedimentary basin and detrital zircon record along East Laurentia and Baltica during assembly and breakup of Rodinia. J. Geol. Soc. 164 , 257–275 (2007).

Strachan, R. A., Olierook, H. K. H. & Kirkland, C. L. Evidence from the U–Pb–Hf signatures of detrital zircons for a Baltican provenance for basal Old Red Sandstone successions, northern Scottish Caledonides. J. Geol. Soc. 178 , https://doi.org/10.1144/jgs2020-241 (2021).

Stevens, T. & Baykal, Y. Detrital zircon U–Pb ages and source of the late Palaeocene Thanet Formation, Kent, SE England. Proc. Geol. Assoc. 132 , 240–248 (2021).

O’Sullivan, G., Chew, D. M., Kenny, G., Heinrichs, I. & Mulligan, D. The trace element composition of apatite and its application to detrital provenance studies. Earth Sci. Rev. 201 , 103044 (2020).

Oliver, G., Wilde, S. & Wan, Y. Geochronology and geodynamics of Scottish granitoids from the late Neoproterozoic break-up of Rodinia to Palaeozoic collision. J. Geol. Soc. 165 , 661–674 (2008).

Fleischer, M. & Altschuler, Z. S. The lanthanides and yttrium in minerals of the apatite group-an analysis of the available data. Neu. Jb. Mineral. Mh. 10 , 467–480 (1986).

Goodenough, K. M., Millar, I., Strachan, R. A., Krabbendam, M. & Evans, J. A. Timing of regional deformation and development of the Moine Thrust Zone in the Scottish Caledonides: constraints from the U–Pb geochronology of alkaline intrusions. J. Geol. Soc. 168 , 99–114 (2011).

Stacey, J. S. & Kramers, J. D. Approximation of terrestrial lead isotope evolution by a two-stage model. Earth Planet. Sci. Lett. 26 , 207–221 (1975).

Evans, J. A. et al. Applying lead (Pb) isotopes to explore mobility in humans and animals. PLoS ONE 17 , e0274831 (2022).

Morton, A., Knox, R. & Frei, D. Heavy mineral and zircon age constraints on provenance of the Sherwood Sandstone Group (Triassic) in the eastern Wessex Basin, UK. Proc. Geol. Assoc. 127 , 514–526 (2016).

Morton, A., Hounslow, M. W. & Frei, D. Heavy-mineral, mineral-chemical and zircon-age constraints on the provenance of Triassic sandstones from the Devon coast, southern Britain. Geologos 19 , 67–85 (2013).

Phillips, E. R., Smith, R. A., Stone, P., Pashley, V. & Horstwood, M. Zircon age constraints on the provenance of Llandovery to Wenlock sandstones from the Midland Valley terrane of the Scottish Caledonides. Scott. J. Geol. 45 , 131–146 (2009).

McKellar, Z., Hartley, A. J., Morton, A. C. & Frei, D. A multidisciplinary approach to sediment provenance analysis of the late Silurian–Devonian Lower Old Red Sandstone succession, northern Midland Valley Basin, Scotland. J. Geol. Soc. 177 , 297–314 (2019).

Beranek, L. P., Gee, D. G. & Fisher, C. M. Detrital zircon U–Pb–Hf isotope signatures of Old Red Sandstone strata constrain the Silurian to Devonian paleogeography, tectonics, and crustal evolution of the Svalbard Caledonides. GSA Bull. 132 , 1987–2003 (2020).

John, B. The Stonehenge Bluestones (Greencroft Books, 2018).

John, B. The Stonehenge bluestones did not come from Waun Mawn in West Wales. The Holocene https://doi.org/10.1177/09596836241236318 (2024).

Clark, C. D. et al. Growth and retreat of the last British–Irish Ice Sheet, 31 000 to 15 000 years ago: the BRITICE-CHRONO reconstruction. Boreas 51 , 699–758 (2022).

Gibbard, P. L. & Clark, C. D. in Developments in Quaternary Sciences , Vol. 15 (eds Ehlers, J. et al.) 75–93 (Elsevier, 2011).

Bevins, R., Ixer, R., Pearce, N., Scourse, J. & Daw, T. Lithological description and provenancing of a collection of bluestones from excavations at Stonehenge by William Hawley in 1924 with implications for the human versus ice transport debate of the monument’s bluestone megaliths. Geoarchaeology 38 , 771–785 (2023).

Snoeck, C. et al. Strontium isotope analysis on cremated human remains from Stonehenge support links with west Wales. Sci. Rep. 8 , 10790 (2018).

Article   ADS   PubMed   PubMed Central   Google Scholar  

Viner, S., Evans, J., Albarella, U. & Pearson, M. P. Cattle mobility in prehistoric Britain: strontium isotope analysis of cattle teeth from Durrington Walls (Wiltshire, Britain). J. Archaeolog. Sci. 37 , 2812–2820 (2010).

Evans, J. A., Chenery, C. A. & Fitzpatrick, A. P. Bronze Age childhood migration of individuals near Stonehenge, revealed by strontium and oxygen isotope tooth enamel analysis. Archaeometry 48 , 309–321 (2006).

Bradley, R. Beyond the bluestones: links between distant monuments in Late Neolithic Britain and Ireland. Antiquity 98 , 821–828 (2024).

Bradley, R. Long distance connections within Britain and Ireland: the evidence of insular rock art. Proc. Prehist. Soc. 89 , 249–271 (2023).

Fairweather, A. D. & Ralston, I. B. M. The Neolithic timber hall at Balbridie, Grampian Region, Scotland: the building, the date, the plant macrofossils. Antiquity 67 , 313–323 (1993).

Bayliss, A., Marshall, P., Richards, C. & Whittle, A. Islands of history: the Late Neolithic timescape of Orkney. Antiquity 91 , 1171–1188 (2017).

Parker Pearson, M. et al. in Megaliths and Geology (eds Bouventura, R. et al.) 151–169 (Archaeopress Publishing, 2020).

Pigière, F. & Smyth, J. First evidence for cattle traction in Middle Neolithic Ireland: A pivotal element for resource exploitation. PLoS ONE 18 , e0279556 (2023).

Article   PubMed   PubMed Central   Google Scholar  

Godwin, H. History of the natural forests of Britain: establishment, dominance and destruction. Philos. Trans. R. Soc. B 271 , 47–67 (1975).

ADS   Google Scholar  

Martínková, N. et al. Divergent evolutionary processes associated with colonization of offshore islands. Mol. Ecol. 22 , 5205–5220 (2013).

Bradley, R. & Edmonds, M. Interpreting the Axe Trade: Production and Exchange in Neolithic Britain (Cambridge Univ. Press, 2005).

Peacock, D., Cutler, L. & Woodward, P. A Neolithic voyage. Int. J. Naut. Archaeol. 39 , 116–124 (2010).

Pinder, A. P., Panter, I., Abbott, G. D. & Keely, B. J. Deterioration of the Hanson Logboat: chemical and imaging assessment with removal of polyethylene glycol conserving agent. Sci. Rep. 7 , 13697 (2017).

Harff, J. et al. in Submerged Landscapes of the European Continental Shelf: Quaternary Paleoenvironments (eds Flemming, N. C. et al.) 11–49 (2017).

Nordsvan, A. R., Kirscher, U., Kirkland, C. L., Barham, M. & Brennan, D. T. Resampling (detrital) zircon age distributions for accurate multidimensional scaling solutions. Earth Sci. Rev. 204 , 103149 (2020).

Ixer, R., Bevins, R. & Turner, P. Alternative Altar Stones? Carbonate-cemented micaceous sandstones from the Stonehenge landscape. Wilts. Archaeol. Nat. Hist. Mag. 112 , 1–13 (2019).

Paton, C., Hellstrom, J. C., Paul, B., Woodhead, J. D. & Hergt, J. M. Iolite: freeware for the visualisation and processing of mass spectrometric data. J. Anal. At. Spectrom. 26 , 2508–2518 (2011).

Vermeesch, P. IsoplotR: a free and open toolbox for geochronology. Geosci. Front. 9 , 1479–1493 (2018).

Jackson, S. E., Pearson, N. J., Griffin, W. L. & Belousova, E. A. The application of laser ablation-inductively coupled plasma-mass spectrometry to in situ U–Pb zircon geochronology. Chem. Geol. 211 , 47–69 (2004).

Sláma, J. et al. Plešovice zircon—A new natural reference material for U–Pb and Hf isotopic microanalysis. Chem. Geol. 249 , 1–35 (2008).

Wiedenbeck, M. et al. Three natural zircon standards for U-Th-Pb, Lu–Hf, trace element and REE analyses. Geostand. Newslett. 19 , 1–23 (1995).

Stern, R. A., Bodorkos, S., Kamo, S. L., Hickman, A. H. & Corfu, F. Measurement of SIMS instrumental mass fractionation of Pb isotopes during zircon dating. Geostand. Geoanal. Res. 33 , 145–168 (2009).

Marsh, J. H., Jørgensen, T. R. C., Petrus, J. A., Hamilton, M. A. & Mole, D. R. U-Pb, trace element, and hafnium isotope composition of the Maniitsoq zircon: a potential new Archean zircon reference material. Goldschmidt Abstr. 2019 , 18 (2019).

Vermeesch, P. On the treatment of discordant detrital zircon U–Pb data. Geochronology 3 , 247–257 (2021).

Gehrels, G. in Tectonics of Sedimentary Basins: Recent Advances (eds Busby, C. & Azor, A.) 45–62 (2011).

Vermeesch, P. How many grains are needed for a provenance study? Earth Planet. Sci. Lett. 224 , 441–451 (2004).

Dröllner, M., Barham, M., Kirkland, C. L. & Ware, B. Every zircon deserves a date: selection bias in detrital geochronology. Geol. Mag. 158 , 1135–1142 (2021).

Zutterkirch, I. C., Kirkland, C. L., Barham, M. & Elders, C. Thin-section detrital zircon geochronology mitigates bias in provenance investigations. J. Geol. Soc. 179 , jgs2021–070 (2021).

Morton, A., Waters, C., Fanning, M., Chisholm, I. & Brettle, M. Origin of Carboniferous sandstones fringing the northern margin of the Wales-Brabant Massif: insights from detrital zircon ages. Geol. J. 50 , 553–574 (2015).

Luvizotto, G. et al. Rutile crystals as potential trace element and isotope mineral standards for microanalysis. Chem. Geol. 261 , 346–369 (2009).

Zack, T. et al. In situ U–Pb rutile dating by LA-ICP-MS: 208 Pb correction and prospects for geological applications. Contrib. Mineral. Petrol. 162 , 515–530 (2011).

Dröllner, M., Barham, M. & Kirkland, C. L. Reorganization of continent-scale sediment routing based on detrital zircon and rutile multi-proxy analysis. Basin Res. 35 , 363–386 (2023).

Liebmann, J., Barham, M. & Kirkland, C. L. Rutile ages and thermometry along a Grenville anorthosite pathway. Geochem. Geophys. Geosyst. 24 , e2022GC010330 (2023).

Zack, T. & Kooijman, E. Petrology and geochronology of rutile. Rev. Mineral. Geochem. 83 , 443–467 (2017).

Thompson, J. et al. Matrix effects in Pb/U measurements during LA-ICP-MS analysis of the mineral apatite. J. Anal. At. Spectrom. 31 , 1206–1215 (2016).

Schmitz, M. D., Bowring, S. A. & Ireland, T. R. Evaluation of Duluth Complex anorthositic series (AS3) zircon as a U–Pb geochronological standard: new high-precision isotope dilution thermal ionization mass spectrometry results. Geochim. Cosmochim. Acta 67 , 3665–3672 (2003).

Schoene, B. & Bowring, S. U–Pb systematics of the McClure Mountain syenite: thermochronological constraints on the age of the 40 Ar/ 39 Ar standard MMhb. Contrib. Mineral. Petrol. 151 , 615–630 (2006).

Thomson, S. N., Gehrels, G. E., Ruiz, J. & Buchwaldt, R. Routine low-damage apatite U–Pb dating using laser ablation-multicollector-ICPMS. Geochem. Geophys. Geosyst. 13 , https://doi.org/10.1029/2011GC003928 (2012).

Barfod, G. H., Krogstad, E. J., Frei, R. & Albarède, F. Lu–Hf and PbSL geochronology of apatites from Proterozoic terranes: a first look at Lu–Hf isotopic closure in metamorphic apatite. Geochim. Cosmochim. Acta 69 , 1847–1859 (2005).

McDowell, F. W., McIntosh, W. C. & Farley, K. A. A precise 40 Ar– 39 Ar reference age for the Durango apatite (U–Th)/He and fission-track dating standard Chem. Geol. 214 , 249–263 (2005).

Kirkland, C. L. et al. Apatite: a U–Pb thermochronometer or geochronometer? Lithos 318-319 , 143–157 (2018).

Simpson, A. et al. In-situ Lu Hf geochronology of garnet, apatite and xenotime by LA ICP MS/MS. Chem. Geol. 577 , 120299 (2021).

Glorie, S. et al. Robust laser ablation Lu–Hf dating of apatite: an empirical evaluation. Geol. Soc. Lond. Spec. Publ. 537 , 165–184 (2024).

Norris, C. & Danyushevsky, L. Towards estimating the complete uncertainty budget of quantified results measured by LA-ICP-MS. Goldschmidt Abstr. 2018 , 1894 (2018).

Nebel, O., Morel, M. L. A. & Vroon, P. Z. Isotope dilution determinations of Lu, Hf, Zr, Ta and W, and Hf isotope compositions of NIST SRM 610 and 612 glass wafers. Geostand. Geoanal. Res. 33 , 487–499 (2009).

Kharkongor, M. B. K. et al. Apatite laser ablation LuHf geochronology: A new tool to date mafic rocks. Chem. Geol. 636 , 121630 (2023).

Glorie, S. et al. Detrital apatite Lu–Hf and U–Pb geochronology applied to the southwestern Siberian margin. Terra Nova 34 , 201–209 (2022).

Spencer, C. J., Kirkland, C. L., Roberts, N. M. W., Evans, N. J. & Liebmann, J. Strategies towards robust interpretations of in situ zircon Lu–Hf isotope analyses. Geosci. Front. 11 , 843–853 (2020).

Jochum, K. P. et al. GeoReM: a new geochemical database for reference materials and isotopic standards. Geostand. Geoanal. Res. 29 , 333–338 (2005).

Janousek, V., Farrow, C. & Erban, V. Interpretation of whole-rock geochemical data in igneous geochemistry: introducing Geochemical Data Toolkit (GCDkit). J. Petrol. 47 , 1255–1259 (2006).

Boynton, W. V. in Developments in Geochemistry , Vol. 2 (ed. Henderson, P.) 63–114 (Elsevier, 1984).

Landing, E., Keppie, J. D., Keppie, D. F., Geyer, G. & Westrop, S. R. Greater Avalonia—latest Ediacaran–Ordovicia “peribaltic” terrane bounded by continental margin prisms (“Ganderia”, Harlech Dome, Meguma): review, tectonic implications, and paleogeography. Earth Sci. Rev. 224 , 103863 (2022).

Download references

Acknowledgements

Funding was provided by an Australian Research Council Discovery Project (DP200101881). Sample material was loaned from the Salisbury Museum and Amgueddfa Cymru–Museum Wales and sampled with permission. The authors thank A. Green for assistance in accessing the Salisbury Museum material; B. McDonald, N. Evans, K. Rankenburg and S. Gilbert for their help during isotopic analysis; and P. Sampaio for assistance with statistical analysis. Instruments in the John de Laeter Centre, Curtin University, were funded via AuScope, the Australian Education Investment Fund, the National Collaborative Research Infrastructure Strategy, and the Australian Government. R.E.B. acknowledges a Leverhulme Trust Emeritus Fellowship.

Author information

Authors and affiliations.

Timescales of Mineral Systems Group, School of Earth and Planetary Sciences, Curtin University, Perth, Western Australia, Australia

Anthony J. I. Clarke & Christopher L. Kirkland

Department of Geography and Earth Sciences, Aberystwyth University, Aberystwyth, UK

Richard E. Bevins & Nick J. G. Pearce

Department of Earth Sciences, The University of Adelaide, Adelaide, South Australia, Australia

Stijn Glorie

Institute of Archaeology, University College London, London, UK

Rob A. Ixer

You can also search for this author in PubMed   Google Scholar

Contributions

A.J.I.C.: writing, original draft, formal analysis, investigation, visualization, project administration, conceptualization and methodology. C.L.K.: supervision, resources, formal analysis, funding acquisition, writing, review and editing, conceptualization and methodology. R.E.B.: writing, review and editing, resources and conceptualization. N.J.G.P.: writing, review and editing, resources and conceptualization. S.G.: resources, formal analysis, funding acquisition, writing, review and editing, supervision, and methodology. R.A.I.: writing, review and editing.

Corresponding author

Correspondence to Anthony J. I. Clarke .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Peer review

Peer review information.

Nature thanks Tim Kinnaird and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer review reports are available.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended data fig. 1 geological maps of potential source terranes for the altar stone..

a , Schematic map of the North Atlantic region with the crystalline terranes in the Caledonian-Variscan orogens depicted prior to the opening of the North Atlantic, adapted after ref.  95 . b , Schematic map of Britain and Ireland, showing outcrops of Old Red Sandstone, basement terranes, and major faults with reference to Stonehenge.

Extended Data Fig. 2 Altar Stone zircon U–Pb data.

a , Tera-Wasserburg plot for all concordant (≤10% discordant) zircon analyses reported from three samples of the Altar Stone. Discordance is defined using the concordia log % distance approach, and analytical ellipses are shown at the two-sigma uncertainty level. The ellipse colour denotes the sample. Replotted isotopic data for thin-section FN593 is from ref. 1 . b , Kernel density estimate for concordia U–Pb ages of concordant zircon from the Altar Stone, using a kernel and histogram bandwidth of 50 Ma. Fifty-six concordant analyses are shown from 113 measurements. A rug plot is given below the kernel density estimate, marking the age of each measurement.

Extended Data Fig. 3 Comparative kernel density estimates of concordant zircon concordia ages from the Altar Stone, crystalline sources terranes, and comparative sedimentary rock successions.

Each plot uses a kernel and histogram bandwidth of 50 Ma. The zircon U–Pb geochronology source for each comparative dataset is shown with their respective kernel density estimate. Zircon age data for basement terranes (right side of the plot) was sourced from refs. 20 , 26 .

Extended Data Fig. 4 Plots of rutile U–Pb ages.

a , Tera-Wasserburg plot of rutile U–Pb analyses from the Altar Stone (thin-section MS3). Isotopic data is shown at the two-sigma uncertainty level. b , Kernel density estimate for Group 2 rutile 207 Pb corrected 206 Pb/ 238 U ages, using a kernel and histogram bandwidth of 25 Ma. The rug plot below the kernel density estimate marks the age for each measurement.

Extended Data Fig. 5 Apatite Tera-Wasserburg U–Pb plots for the Altar Stone and Orcadian Basin.

a , Altar Stone apatite U–Pb analyses from thin-section MS3. b , Orcadian Basin apatite U–Pb analyses from sample AQ1, Spittal, Caithness. c , Orcadian Basin apatite U–Pb analyses from sample CQ1, Cruaday, Orkney. All data are shown as ellipses at the two-sigma uncertainty level. Regressions through U–Pb data are unanchored.

Extended Data Fig. 6 Combined kernel density estimate and histogram for apatite Lu–Hf single-grain ages from the Altar Stone.

Lu–Hf apparent ages from thin-section 2010K.240. Kernel and histogram bandwidth of 50 Ma. The rug plot below the kernel density estimate marks each calculated age. Single spot ages are calculated assuming an initial average terrestrial 177 Hf/ 176 Hf composition (see  Methods ).

Extended Data Fig. 7 Apatite trace element classification plots for the Altar Stone thin-section MS3.

Colours for all plots follow the geochemical discrimination defined in A. a , Reference 29  classification plot for apatite with an inset pie chart depicting the compositional groupings based on these geochemical ratios. b , The principal component plot of geochemical data from apatite shows the main eigenvectors of geochemical dispersion, highlighting enhanced Nd and La in the distinguishing groups. Medians for each group are denoted with a cross. c , Plot of total rare earth elements (REE) (%) versus (Ce/Yb) n with Mahalanobis ellipses around compositional classification centroids. A P = 0.5 in Mahalanobis distance analysis represents a two-sided probability, indicating that 50% of the probability mass of the chi-squared distribution for that compositional grouping is contained within the ellipse. This probability is calculated based on the cumulative distribution function of the chi-squared distribution. d , Chondrite normalized REE plot of median apatite values for each defined apatite classification type.

Extended Data Fig. 8 Cumulative probability density function plot.

Cumulative probability density function plot of comparative Old Red Sandstone detrital zircon U–Pb datasets (concordant ages) versus the Altar Stone. Proximity between cumulative density probability lines implies similar detrital zircon age populations.

Supplementary information

Supplementary information 1.

Zircon, rutile, and apatite U–Pb data for the Altar Stone and Orcadian Basin samples. A ) Zircon U–Pb data for MS3, 2010K.240, and FN593. B ) Zircon U–Pb data for secondary references. C ) Rutile U–Pb data for MS3. D ) Rutile U–Pb data for secondary references. E ) Session 1 apatite U–Pb data for MS3. F ) Session 1 apatite U–Pb data for secondary references. G ) Session 2 apatite U–Pb data for Orcadian Basin samples. H ) Session 2 apatite U–Pb data for secondary references.

Reporting Summary

Peer review file, supplementary information 2.

Apatite Lu–Hf data for the Altar Stone. A) Apatite Lu–Hf isotopic data and ages for thin-section 2010K.240. B) Apatite Lu–Hf data for secondary references.

Supplementary Information 3

Apatite trace elements for the Altar Stone. A) Apatite trace element data for MS3. B) Apatite trace element secondary reference values.

Supplementary Information 4–8

Supplementary Information 4 : Summary of analyses. Summary table of analyses undertaken in this work on samples from the Altar Stone and the Orcadian Basin. Supplementary Information 5: Summary of zircon U–Pb reference material. A summary table of analyses was obtained for zircon U–Pb secondary reference material run during this work. Supplementary Information 6: Kolmogorov–Smirnov test results. Table of D and P values for the Kolmogorov–Smirnov test on zircon ages from the Altar Stone and potential source regions. Supplementary Information 7: Kolmogorov–Smirnov test results, with Monte Carlo resampling. Table of D and P values for the Kolmogorov–Smirnov test (with Monte Carlo resampling) on zircon ages from the Altar Stone and potential source regions. Supplementary Information 8: Summary of apatite U–Pb reference material. A summary table of analyses was obtained for the apatite U–Pb secondary reference material run during this work.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Clarke, A.J.I., Kirkland, C.L., Bevins, R.E. et al. A Scottish provenance for the Altar Stone of Stonehenge. Nature 632 , 570–575 (2024). https://doi.org/10.1038/s41586-024-07652-1

Download citation

Received : 16 December 2023

Accepted : 03 June 2024

Published : 14 August 2024

Issue Date : 15 August 2024

DOI : https://doi.org/10.1038/s41586-024-07652-1

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Stonehenge’s enigmatic centre stone was hauled 800 kilometres from scotland.

  • Nisha Gaind

Nature (2024)

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

research about chemical weathering

IMAGES

  1. Chemical Weathering

    research about chemical weathering

  2. Chemical Weathering Process, Examples, Types & Diagram

    research about chemical weathering

  3. Chemical Weathering: A Great Natural Force

    research about chemical weathering

  4. 4 Types and Examples of Chemical Weathering

    research about chemical weathering

  5. 4 Types and Examples of Chemical Weathering

    research about chemical weathering

  6. PPT

    research about chemical weathering

COMMENTS

  1. Chemical Weathering

    Chemical weathering of rocks is a spontaneous (i.e., irreversible) thermodynamic process leading to a more stable state for natural materials under a given set of conditions (e.g., temperature and pressure). It results from the reaction of aqueous, acidic, and oxidizing solutions with the minerals in rocks and soils.

  2. The contribution of living organisms to rock weathering in the critical

    The (bio)chemical weathering of rocks is an important consequence of these forcings, along with other processes such as erosion. ... future research should continue to identify the "actors ...

  3. Enhanced chemical weathering as a geoengineering strategy to reduce

    [1] Chemical weathering is an integral part of both the rock and carbon cycles and is being affected by changes in land use, particularly as a result of agricultural practices such as tilling, mineral fertilization, or liming to adjust soil pH. These human activities have already altered the terrestrial chemical cycles and land-ocean flux of major elements, although the extent remains ...

  4. Global chemical weathering dominated by continental arcs since ...

    Earth's plate-tectonic activity regulates the carbon cycle and, hence, climate, via volcanic outgassing and silicate-rock weathering. Mountain building, arc-continent collisions and clustering ...

  5. The chemical, mechanical, and hydrological evolution of weathering

    1 Introduction. Chemical weathering impacts the Earth's surface by providing mineral-derived nutrients to ecosystems [Chadwick et al., 1999], releasing solutes to streams and oceans [Stallard and Edmond, 1983], producing sediment for transport [Riebe et al., 2015], and regulating the Earth's climate through feedbacks on the atmospheric concentration of CO 2 [Berner et al., 1983; Hilley et al ...

  6. Globally elevated chemical weathering rates beneath glaciers

    Metrics. Physical erosion and chemical weathering rates beneath glaciers are expected to increase in a warming climate with enhanced melting but are poorly constrained. We present a global dataset ...

  7. Chemical Weathering

    Laboratory research on chemical weathering has a long history (about a century), and the nature of these involved processes is still undergoing debate within the scientific community. Most experiments in dissolution yield chemical weathering rates up to four orders of magnitude greater than those estimated from field studies.

  8. Cation isotopes trace chemical weathering

    1. Introduction. Chemical weathering is the processes whereby crustal minerals are hydrolyzed (and possibly oxidized) by soil solutions and, therefore, is the primary mechanism by which streams, rivers, and ultimately the oceans acquire dissolved salts [1].It alters the geochemical compositions of the lithosphere, hydrosphere, and atmosphere, regulates global nutrient and carbon cycles, and ...

  9. Chemical Weathering

    Research into the degradation and conservation of the "natural stone" of historic monuments has also recently been a great incentive for studies about weathering. In nature, the chemical weathering of rocks and soils rarely works alone; in fact, it is closely associated with physical weathering, often being initiated and potentiated by the ...

  10. Chemical Weathering

    Chemical Weathering is an erosional process in which rocks and other materials are broken down by chemical reactions (predominantly by water and chemicals dissolved within it). A. Types of Chemical Weathering. 1. dissolution. This process occurs when water comes into contact with rocks and dissolves the minerals that make up that rock into ...

  11. Enhanced chemical weathering: A solution to the climate crisis?

    High temperatures accelerate the chemical rock weathering process, reducing the levels of carbon dioxide in the atmosphere, allowing the climate to recover," said Pogge von Strandmann.

  12. Enhanced chemical weathering: A solution to the climate crisis?

    High temperatures accelerate the chemical rock weathering process, reducing the levels of carbon dioxide in the atmosphere, allowing the climate to recover," said Pogge von Strandmann. Climate ...

  13. Chemical Weathering, Atmospheric CO2, and Climate

    There has been considerable controversy concerning the role of chemical weathering in the regulation of the atmospheric partial pressure of carbon dioxide, and thus the strength of the greenhouse effect and global climate. Arguments center on the sensitivity of chemical weathering to climatic factors, especially temperature. Laboratory studies reveal a strong dependence of mineral dissolution ...

  14. Chemical weathering

    Research chevron_right … Earth surface geochemistry chevron_right; Chemical weathering; Chemical weathering. Chemical weathering is a key Earth surface process, one that shapes the landscape, supplies dissolved constituents to the oceans and, on geological timescales, controls atmospheric CO 2 concentrations. Childpage navigation ...

  15. What Is Chemical Weathering?

    Chemical weathering can change the composition and shape of rocks. Oxidation turned this peridotite to different tones of red-rusty brown. There are three types of weathering which affect rock: physical, biological, and chemical. Chemical weathering, also known as decomposition or decay, is the breakdown of rock by chemical mechanisms.

  16. Weathering

    Weathering describes the breaking down or dissolving of rocks and minerals on the surface of Earth. Water, ice, acids, salts, plants, animals, and changes in temperature are all agents of weathering. Once a rock has been broken down, a process called erosion transports the bits of rock and mineral away. No rock on Earth is hard enough to resist the forces of weathering and erosion.

  17. Enhanced weathering strategies for stabilizing climate and ...

    Chemical breakdown of rocks, weathering, is an important but very slow part of the carbon cycle that ultimately leads to CO2 being locked up in carbonates on the ocean floor. ... Research suggests ...

  18. FSU Research: Chemical weathering could alleviate some climate change

    A team of Florida State University scientists has discovered that chemical weathering, a process in which carbon dioxide breaks down rocks and then gets trapped in sediment, can happen at a much faster rate than scientists previously assumed and could potentially counteract some of the current and future climate change caused by humans.

  19. Weathering

    Weathering is the deterioration of rocks, soils and minerals (as well as wood and artificial materials) through contact with water, atmospheric gases, sunlight, and biological organisms.It occurs in situ (on-site, with little or no movement), and so is distinct from erosion, which involves the transport of rocks and minerals by agents such as water, ice, snow, wind, waves and gravity.

  20. (PDF) Weathering, Erosion, and Susceptibility to Weathering

    Weathering is the chemical and physical change in time. of ground under influence of atmosphere, hydrosphere, cryosphere, biosphere, and nuclear radiation (temperature, rain, circulating ...

  21. Graphic representation of chemical weathering

    The several graphic devices available for use with studies of chemical weathering are analytically reviewed. The specific advantages and disadvantages of each are noted. A modification of existing methods, combining a number of their more desirable aspects, is proposed and illustrated. This content is PDF only.

  22. Resolving controversies surrounding carbon sinks from carbonate weathering

    The importance of carbonate weathering carbon sinks (CCSs) is almost equal to that of vegetation photosynthesis in the global carbon cycle. However, CCSs have become controversial in formulating carbon neutral policies to deal with global climate problems in various countries, since the carbonate dissolution is reversible. In order to address these controversies, we reviewed recent advances in ...

  23. Design of a Weathering Chamber for UV Aging of Microplastics in the

    Microplastics are an ever-growing concern in the environment. Their degradation may lead to greater absorption of toxic pollutants, which may ultimately pose a threat to human health. In the pursuit of understanding microplastics' fate, behavior, and toxicity, there is a vital need to understand their aging and weathering. For this, multiple weathering setup designs were put forward. However ...

  24. Correlations of chemical weathering indicators with major ...

    During chemical weathering, there are significant differences in the geochemical behaviours of Rb and Sr, and the ratio of the two is a good geochemical indicator of the degree of chemical ...

  25. Weathering Scratches the Surface of Plastic

    Plastics might be more resilient than you think—new research shows they can survive in the ocean for decades. A study released in Environmental Science & Technology documents how commercially available macroplastics (>5mm) can withstand decades of exposure in marine environments with minimal degradation. With the influx of plastic waste into the world's oceans, it is imperative to ...

  26. Researcher Insights: Environmental Weathering Transforms Plastic

    In addition to influencing fate, weathering-induced transformations can influence the ecotoxicological impact, or hazard, associated with exposure. Recently, we have observed that the same starting material - in this case, polystyrene - will impact sewage sludge in an anaerobic digestor differently based on the path of weathering.

  27. Potential for large-scale CO2 removal via enhanced rock weathering with

    Enhanced silicate rock weathering (ERW), deployable with croplands, has potential use for atmospheric carbon dioxide (CO2) removal (CDR), which is now necessary to mitigate anthropogenic climate ...

  28. For Stonehenge's Altar Stone, an Improbably Long Ancient Journey

    The researchers analyzed the chemical composition and the ages of mineral grains in two microscopic fragments of the Altar Stone. ... chemically stable minerals that are highly resistant to ...

  29. A Scottish provenance for the Altar Stone of Stonehenge

    Mineral ages and chemical analysis of fragments of the Altar Stone from the Neolithic stone circle at Stonehenge suggest that it was transported from northeast Scotland, more than 750 km away ...