Get Your ALL ACCESS Shop Pass here →
Beach Erosion Project
Have you ever noticed what happens to the coast line when a big storm rolls through? Where did the beach go? What you are noticing is the effect of coastal erosion, and now you can set up a beach erosion demonstration to show your kids what’s happening. This fun and easy ocean science activity is sure to be a hit with your kids, with hands-on learning!
Explore Erosion For Earth Science
Break out the sensory play as you prepare to add this beach erosion activity to your ocean-theme lesson plans. If you want to learn about what’s happening between the sand and the waves, let’s dig in (literally!). While you’re at it, check out more fun ocean activities , experiments, and ocean crafts.
Our earth science activities are designed with you, the parent or teacher, in mind! They are easy to set up and quick to do. Most activities will take only 15 to 30 minutes to complete and are heaps of fun! Plus, our supplies lists usually contain only free or cheap materials you can source from home!
Let’s explore beach erosion by building a model! This is a great hands-on ocean STEM activity that is sure to get kids thinking!
What Is Beach Erosion?
Beach erosion is the loss of beach sand, usually from a combination of wind and water movement such as waves and currents. Sand is moved off the beach or shore by these things and is transferred to deeper water.
This process makes beaches appear shorter and lower. You can see severe beach erosion after a strong storm like a hurricane.
TRY: Learn more about erosion with an edible soil layer model and this fun soil erosion activity.
You’ll also be sent weekly projects to your inbox! We respect your privacy. Unsubscribe at any time.
Erosion Experiment
Dive into Earth Science and explore what happens to our beaches due to coastal erosion.
Watch the Video
- White paint pan
- Blue food coloring
- Plastic bottle
- Large pan or tray.
How To Set Up A Beach Erosion Model
STEP 1: Add about 5 cups of sand to one side of your pan. You will want to build it up on a slope so that when water is added some of the sand is higher.
STEP 2: Place some rocks or shells in the sand for a beach theme!
STEP 3: Fill a small bottle with water, add a drop of blue food coloring, shake and pour into the deep part of your pan.
STEP 4: Add 4 more cups of water.
STEP 5: Use the empty bottle to press up and down in the water to make waves.
STEP 6: Pay attention to how the water affects the sand. What happens if the waves move faster or slower?
How Can We Stop Coastal Erosion?
Coastal erosion is the loss of coastal land due to the removal of sand or rock from the shoreline. Sadly, building along the coast can damage sand dunes.
Dunes are mounds of sand that separate the beach you walk on and higher ground. The roots of dune grasses help keep the sand in place. Try not to walk on the dune grasses, so they are not destroyed!
People sometimes build walls called jetties that stick out into the ocean and change the movement of the sand.
Seawalls can also help with erosion. They are structures that separate land and water and generally help prevent erosion from large waves. Seawalls are more significant structures where flooding is more common. Please don’t remove rocks from the seawall!
Classroom Tips
This beach erosion activity asks a few questions!
- What is coastal erosion?
- What causes beach erosion?
- How can we stop erosion?
Let’s explore the answers together!
Be prepared! Kids are going to loving play with this, and it could get a bit messy!
Further Extension: Have kids develop ideas for something they can make that will help prevent beach erosion during a storm!
More Ocean Experiments For Kids
- Oil Spill Cleanup Experiment
- Layers of the Ocean
- How Do Whales Stay Warm?
- Ocean Waves In A Bottle
- Ocean Acidification: Seashells In Vinegar Experiment
- Fun Facts About Narwhals
- Ocean Currents Activity
Printable Ocean Activities Pack
If you want to have all of your printable ocean activities in one convenient place, plus exclusive worksheets with an ocean theme, our 100+ page Ocean STEM Project Pack is what you need!
Check out The Complete Ocean Science and STEM Pack in our SHOP!
Subscribe to receive a free 5-Day STEM Challenge Guide
~ projects to try now ~.
Science Fun
Coastal Erosion Ocean Science Experiment
In this fun and easy ocean science experiment, we’re going to explore and investigate coastal erosion.
Instructions:
- Pile sand on one end of the baking pan and firmly pat it down to create a gradually sloping beach.
- Pour some water into the pan until the sand is partly covered.
- First gently and then progressively faster slide the pan back and forth to create waves that was up on your beach.
- Observe the sand move and shift.
EXPLORE AWESOME SCIENCE EXPERIMENT VIDEOS!
How it Works:
Coastal erosion is the loss or displacement of land from processes such as waves and other actions that remove sand, sediment, and bedrock. This ocean experiment allows you to visualize coastal erosion. This experiment uses water to create erosion but wind is also another contributor to coastal erosion. Storms, currents, and even ice are other contributors to coastal erosion.
Make This A Science Project:
Try filling little sandwich baggies with sand and place these sand bags between the pile of sand and water.
EXPLORE TONS OF FUN AND EASY SCIENCE EXPERIMENTS!
SUBSCRIBE AND NEVER MISS A NEW SCIENCE FUN VIDEO!
previous experiment
Next experiment.
Choose an Account to Log In
Notifications
Science project, erosion experiment.
Huff, puff and blow as hard as you can—do you think you can blow down a sand castle with your mighty breath? Blowing air erodes —wears away—sand and dirt. In Three Little Pigs , houses of straw or twigs were no match for the Big Bad Wolf's strong lungs. Do you think a house made of sand or mud could hold up to wind or water? Let's find out in this erosion experiment.
In some parts of the world, houses are still made from mud and sand, which makes them easier to knock down than homes made of brick and stone. To understand how these materials can (and should) be used to make homes, you'll need to explore how they stand up to the elements.
How does erosion affect structures made of sand and soil?
- Garden hose
- Plastic cups
- Battery-operated portable fan
- Stopwatch (optional)
- Before you start your erosion experiment, take a look at your ingredients. How does the sand feel? How does the soil feel? Do sand and soil stick together easily, or fall apart? Record any observations —things you see—in your notebook.
- Look over your notes, and make a guess about what you think will happen to your soil structure, and what will happen to your sand structure, when you try to knock them down with wind or water. Which "castle" will last longer? Write your guess—called a hypothesis —in your notebook.
- Take your ingredients to a hard surface, like concrete or asphalt. These smooth surfaces will allow you to see exactly how the water and wind affect your structures.
- Create mud by mixing some soil with water. Dirt and soil may look the same, but technically they aren't. Soil consists of sand, silt, clay, minerals, water, air and organic material. And you thought it was just dirt, didn't you?
- Create wet sand by mixing sand with just enough water to make it stick together—like sand on the beach that you use to make sand castles.
- Use a plastic cup to make a structure out of the wet soil; maybe it'll be a castle, maybe just a lot of cups turned over in your own design. Be creative and make your structures just the way YOU like them.
- Repeat Step 6 with the wet sand.
- Next, make similar structures with dry soil and dry sand, for a total of four structures: one made from muddy soil, one dry soil, one wet sand, and one dry sand.
- Now that your creations exist, it's time to see which stand up to wind and water. Use your battery-operated fan to blow air on the four structures. Remember, do NOT use an electric fan—mixing water and electricity is dangerous, so make sure you have the right fan before starting.
- What happens to each of your structures? Draw a picture or write what happened in your notebook.
- Take a garden hose and drop it about two feet away from your structures. Turn on the water for about 30 seconds, and then turn it off. Draw a picture or write what happened in your notebook.
The wind from the fan should have blown away the individual particles of your dry sand structure easily. While all of the structures probably ended up washed away by wind or water, the structures made with soil should have been stronger than the structures made with sand.
Dry sand is like dry sugar. It piles up, but wind and water wash it away quickly, because the individual grains of sand don't stick to each other. Wet sand sticks together when combined with just the right amount of water. Too much water washes it away, but it takes more effort than the dry sand because water fills in the space between the grains enough to make the grains stick to each other. You will find that it takes a fan on higher settings to erode, or gradually wear away the sand.
Dry soil holds together better than dry sand. When it's wet, the organic matter and minerals fill the gaps in the soil to hold it tightly together. It takes a strong force (push) of water and wind to knock down a soil castle. This is why ancient Egyptians added straw to their mud when they made bricks for their homes. They molded the mud and straw, allowing them to dry in the sun, becoming bricks. Bricks are hard. Structures, like homes built today with bricks, are harder to tumble. Remember what happened in the Three Little Pigs ?
Now that you've learned about sand and soil with this erosion experiment, keep the science going by testing more elements. Make this project more challenging by recording the time it takes for the structures to erode away with a stopwatch. What would happen if both wind and water hit the castles at the same time? What if you placed your hose at an angle or held it in the air and sprinkled in on your castles? Being a scientist is all about guessing and testing!
Related learning resources
Add to collection, create new collection, new collection, new collection>, sign up to start collecting.
Bookmark this to easily find it later. Then send your curated collection to your children, or put together your own custom lesson plan.
June 21, 2012
Shoreline Science: Exploring the Erosive Energy of Waves
A sandy science activity from Science Buddies
By Science Buddies
Key concepts Oceans Beaches Geology Erosion Introduction A day at the beach is a wonderful way to spend time with your family and friends. You can swim, play games and build sand castles. But have you ever wondered how the beach you are standing on came to be? How, for example, did all of that sand get there? Beaches are formed and continually changed by the ocean's waves moving rock particles onshore, offshore and along the shore. In this activity, you can investigate how beach formations are made by some parts of a beach that can resist erosion from the waves more than other parts. Background A beach is a geologic formation made up of loose rock particles such as sand, gravel and shell fragments deposited along the shoreline of a body of water. A beach has a few key features. The berm is the part that is mostly above water; this is the active shoreline. The top of the berm is known as the crest, and the part that slopes toward the water is called the face. At the bottom of the face there may be a trough and, further seaward, there may be sandbars parallel to the beach. The erosion of rock formations in the water, coral reefs and headlands create rock particles that the waves move onshore, offshore and along the shore, creating the beach. Continual erosion of the shoreline by waves also changes the beach over time. One change that erosion can cause is the appearance of a headland. This is land that juts out from the coastline and into the water and affects how the surrounding shoreline is eroded. Materials • Paint-roller pan • Measuring cup • Sand • Water • Timer • Digital camera • Plastic 500-milliliter water bottle (empty) • Adult volunteer to help take pictures • Small gravel, such as aquarium gravel Preparation • Cover the bottom of the paint-roller pan with five cups of sand. Build up a beach with most, but not all, of the sand at the shallow end of the pan. • Slowly pour six cups of water into the deep end of the pan. Let the water and sand settle for five minutes. How has the beach changed during this time? Procedure • Take a picture of your beach so that you have a record of how it looked in its original state. Where is the shoreline (the area where beach and water meet)?* • Lay a plastic bottle horizontally so it is floating in the water in the deep end of the pan. • For two minutes bob the water bottle up and down with your fingertips to create waves. If the waves get so big that water splashes out of the pan, make them smaller. How does the water swirl? How does the shoreline change after one minute? What about after two minutes? • After two minutes of bobbing the bottle, take a picture of the beach. How does it look compared with the first picture? • Empty, clean and dry the paint-roller pan. Prepare a "beach" again, as you did for the preparation. When the beach is complete, make a "headland" by creating a mound out of two cups of small gravel in the middle of the shoreline. The headland should be partly in the water and partly on the beach. Take a picture of the beach with the headland. • Again, lay the plastic bottle horizontally so it is floating in the water. For two minutes, bob the water bottle up and down with your fingertips. Again, if the waves are so big that water splashes out, make them smaller. How does the water swirl? How does the shoreline change after one minute? What about after two minutes? • After two minutes, take a picture of the beach. How does it look compared with the previous picture? • How does the headland affect where the water goes? How does it affect how much the shoreline erodes? • Extra : Repeat this activity at least two more times with a ruler taped to the side of the pan. Exactly how much shoreline erosion occurs with and without a headland? • Extra: Try increasing or decreasing the speed of bobbing the bottle. Does this affect how the beach changes over time? • Extra: Pour a large volume of water all at once into the deep end of the pan to simulate a storm surge or a tsunami. What happens to the beach?
* Correction (6/26/12): The sentence was edited after posting to clarify meaning. In this activity, experimenters are asked to observe changes in the shoreline.
Observations and results Did the shoreline erode, or recede from the water, after you bobbed the water bottle up and down for two minutes? Did most of the shoreline erode less when there was a headland, especially the shoreline closest to it? As waves hit the shoreline over time they erode it and push it further inland. When larger and stronger waves hit the shoreline, such as in a storm, more shoreline is eroded. On a beach that is made up of a mixture of small sand grains and larger, dense rocks, the sand will be eroded away first, leaving behind the larger rocks. Over time this can create a headland—an outcropping of the larger rocks—and a bay nearby. The headland receives most of the waves' energy and consequently protects the bay from erosion. Artificial headlands are sometimes created for this purpose: to prevent coastal erosion. In your model, you should have seen that less shoreline eroded near the headland, but further away from the headland, along the sides of the pan, more erosion occurred and the shoreline was pushed farther inland because the more distant shoreline was not as well protected by the headland. More to explore What Causes Beach Erosion? from Scientific American Shoreline erosion and migration from the State of Delaware A guide to managing coastal erosion in beach/dune systems: Artificial headlands from the Scottish Natural Heritage Restore Your Shore from the Minnesota Department of Natural Resources Building Beaches from Science Buddies
On supporting science journalism
If you're enjoying this article, consider supporting our award-winning journalism by subscribing . By purchasing a subscription you are helping to ensure the future of impactful stories about the discoveries and ideas shaping our world today.
This activity brought to you in partnership with Science Buddies
Soil Erosion Experiment - Teach Kids How Soil Erosion Occurs
Posted by Admin / in Physics Experiments
Erosion occurs when water washes away dirt, rock, or sand. Erosion happens everyday during man-made events, rainfall, and in the ocean. You have probably seen how people try to defend against erosion by placing large rocks near the edge of a river, lake, or the ocean to try to stop water from washing away the land. Erosion can also be a problem for farmers. When a farmer plows their field, there is almost no protection against erosion. Rainfall can easily wash away valuable topsoil which helps the farmer grow things. The soil that is washed away from the farmer's fields can also cause problems for others. The soil sediment can cause ditches, streams, and rivers to become filled. This can result in slower drainage and flooding people's land and buildings. So erosion can be a problem for us and for the fish and other creatures that live in the water, but let's see how it works.
Items Needed for Experiment
- Soil or sand
- Empty milk jug
- Small piece of clay
EXPERIMENT STEPS
Step 1: Place the board on the ground. This experiment is best performed outside because it will result in soil washing off the board.
Step 2: Place a thin layer of soil over the entire board. The soil should be about 1 inch deep for this experiment.
Step 3: Tilt up the board so it is not level. One end of the board should be 3 to 4 inches (75 to 100 mm) higher than the other end. Use some of the extra soil or some rocks under one end to prop it up.
Step 4: Cut a 1/2 inch (13 mm) diameter hole in the bottom of the milk jug using the scissors. The hole does not need to be perfectly round, but the size should be close to 1/2 inch (13 mm).
Step 5: Plug the hole in the milk jug using a small piece of clay so the jug holds water.
Step 6: Fill the milk jug with water.
Step 7: Place the milk jug filled with water on the high side of the board.
Step 8: Release the water by removing the piece of clay from the opening in the bottom of the jug.
Step 9: Observe what happens to the soil on the board.
SCIENCE LEARNED
Soil erosion has become a problem in most areas of the world where there is development. Soil washes into streams and rivers which then exits into lakes and oceans. Water quality is impacted by the amount of sediment washed into the water. Waterways with boat traffic also require dredging to remove the excess sediment in navigation channels in rivers and lakes to keep boats from hitting the bottom. Farm land is also impacted when valuable topsoil is lost to erosion.
There are many ways to slow down soil erosion including:
1. Using silt fences
2. Allowing trees and plants to grow also the side of rivers, streams, lakes, and seas
3. Control the path of water flow and slow down the velocity of the water by re-sloping the ground and ditches which allow runoff
4. Keeping plant growth on the ground also helps keep the soil in place because of the plant roots.
5. Where water travels too quickly, large stones are used to stop erosion.
- About the author
- Back to Experiment
Please select the social network you want to share this page with:
We like you too :)
Thanks for taking time to give us feedback!
- physics experiments
- science experiments for kids
- soil erosion experiment
- water quality experiment
- teach kids about soil erosion
Douglas Rogers - Douglas is a professional engineer and supports math and science education for kids.
- previous experiment
- next experiment
Compass Experiment
in Physics Experiments
Learn how to build a simple wet compass and dry compass, then compare the two types.
How to Calculate Buoyancy Force
Experiment to determine the buoyancy forces acting on a floating and submerged object.
How to Make a Balloon Rocket
Use a balloon to propel a rocket across the sky. Teaches kids simple physics concepts.
How to Make a Paper Airplane Glider
Learn how to build a simple paper airplane glider.
Scientists Develop “Shockingly” Simple Solution To Combat Coastal Erosion
A new method developed by Northwestern University uses electrical currents to solidify marine sand, creating durable, rock-like structures that could replace costly traditional coastal defenses like sea walls.
Researchers from Northwestern University have demonstrated that a zap of electricity can strengthen a marine coastline for generations, mitigating the rising threat of erosion in the face of rising sea levels and climate change.
In their study, recently published in Communications Earth and the Environment , the researchers took inspiration from clams, mussels, and other shell-dwelling sea life, which use dissolved minerals in seawater to build their shells. Similarly, the researchers leveraged the same naturally occurring, dissolved minerals to form a natural cement between sea-soaked grains of sand. But, instead of using metabolic energy like mollusks do, the researchers used electrical energy to spur the chemical reaction.
In laboratory experiments, a mild electrical current instantaneously changed the structure of marine sand, transforming it into a rock-like, immovable solid. The researchers are hopeful this strategy could offer a lasting, inexpensive, and sustainable solution for strengthening global coastlines.
“Over 40% of the world’s population lives in coastal areas. Because of climate change and sea-level rise, erosion is an enormous threat to these communities. Through the disintegration of infrastructure and loss of land, erosion causes billions of dollars in damage per year worldwide. Current approaches to mitigate erosion involve building protection structures or injecting external binders into the subsurface,” said Alessandro Rotta Loria, Louis Berger Assistant Professor of Civil and Environmental Engineering at Northwestern’s McCormick School of Engineering, who led the study.
“My aim was to develop an approach capable of changing the status quo in coastal protection — one that didn’t require the construction of protection structures and could cement marine substrates without using actual cement. By applying a mild electric stimulation to marine soils, we systematically and mechanistically proved that it is possible to cement them by turning naturally dissolved minerals in seawater into solid mineral binders — a natural cement.”
Challenges in Current Coastal Defense Strategies
From intensifying rainstorms to rising sea levels, climate change has created conditions that are gradually eroding coastlines. According to a 2020 study by the European Commission’s Joint Research Centre, nearly 26% of the Earth’s beaches will be washed away by the end of this century.
To mitigate this issue, communities have implemented two main approaches: building protection structures and barriers, such as sea walls, or injecting cement into the ground to strengthen marine substrates, widely consisting of sand. However, multiple problems accompany these strategies. Not only are these conventional methods extremely expensive, but they also do not last.
“Sea walls, too, suffer from erosion,” Rotta Loria said. “So, over time, the sand beneath these walls erodes, and the walls can eventually collapse. Oftentimes, protection structures are made of big stones, which cost millions of dollars per mile. However, the sand beneath them can essentially liquify because of a number of environmental stressors, and these big rocks are swallowed by the ground beneath them.
“Injecting cement and other binders into the ground has a number of irreversible environmental drawbacks. It also typically requires high pressures and significant interconnected amounts of energy.”
Eco-Friendly Electrocementation Process
To bypass these issues, Rotta Loria and his team developed a simpler technique, inspired by coral and mollusks. Seawater naturally contains a myriad of ions and dissolved minerals. When a mild electrical current (2 to 3 volts) is applied to the water, it triggers chemical reactions. This converts some of these constituents into solid calcium carbonate — the same mineral mollusks use to build their shells. Likewise, with a slightly higher voltage (4 volts), these constituents can be predominantly converted into magnesium hydroxide and hydromagnesite, a ubiquitous mineral found in various stones.
When these minerals coalesce in the presence of sand, they act like glue, binding the sand particles together. In the laboratory, the process also worked with all types of sands — from common silica and calcareous sands to iron sands, which are often found near volcanoes.
“After being treated, the sand looks like a rock,” Rotta Loria said. “It is still and solid, instead of granular and incohesive. The minerals themselves are much stronger than concrete, so the resulting sand could become as strong and solid as a sea wall.”
While the minerals form instantaneously after the current is applied, longer electric stimulations garner more substantial results. “We have noticed remarkable outcomes from just a few days of stimulations,” Rotta Loria said. “Then, the treated sand should stay in place, without needing further interventions.”
Sustainable Applications and Future Prospects
Rotta Loria predicts the treated sand should keep its durability, protecting coastlines and property for decades. He also says there is no need to worry about negative effects on sea life. The voltages used in the process are too mild to feel. Other researchers have used similar processes to strengthen undersea structures or even restore coral reefs. In those scenarios, no sea critters were harmed.
If communities decide they no longer want the solidified sand, Rotta Loria has a solution for that, too, as the process is completely reversible. When the battery’s anode and cathode electrodes are switched, the electricity dissolves the minerals — effectively undoing the process.
“The minerals form because we are locally raising the pH of the seawater around cathodic interfaces,” Rotta Loria said. “If you switch the anode with the cathode, then localized reductions in pH are involved, which dissolve the previously precipitated minerals.”
The process offers an inexpensive alternative to conventional methods. After crunching the numbers, Rotta Loria’s team estimates that his process costs just $3 to $6 per cubic meter of electrically cemented ground. More established, comparable methods, which use binders to adhere and strengthen sand, cost up to $70 for the same unit volume.
Research in Rotta Loria’s lab shows this approach can also heal cracked reinforced concrete structures. Much of the existing shoreside infrastructure is made of reinforced concrete, which disintegrates due to complex effects caused by sea-level rise, erosion, and extreme weather. If these structures crack, the new approach bypasses the need to rebuild the infrastructure fully. Instead, one pulse of electricity can heal potentially destructive cracks.
“The applications of this approach are countless,” Rotta Loria said. “We can use it to strengthen the seabed beneath sea walls or stabilize sand dunes and retain unstable soil slopes. We could also use it to strengthen protection structures, marine foundations, and so many other things. There are many ways to apply this to protect coastal areas.”
Next, Rotta Loria’s team plans to test the technique outside of the laboratory and on the beach.
Reference: “Electrodeposition of calcareous cement from seawater in marine silica sands” by Andony Landivar Macias, Steven D. Jacobsen and Alessandro F. Rotta Loria, 22 August 2024, Communications Earth & Environment . DOI: 10.1038/s43247-024-01604-3
The study was supported by the Army Research Office (grant number W911NF2210291) and Northwestern’s Center for Engineering Sustainability and Resilience.
Earth Science Week Classroom Activities
Erosion in a bottle, activity source:.
Source: Soil Science Society of America. Adapted with permission.
Soil erosion is the process of moving soil by water or wind — this happens naturally or through human interference. Preventing soil erosion is important because nutrients are lost, and sediment that accumulates in waterways impacts life there. Conserving soil depends on how it is protected by plants and coverings. You will model erosion by water and compare the amounts of runoff and soil loss generated from three different ground cover types.
• Photos showing erosion • 3 plastic 2-liter bottles • Dry soil (enough to fill each bottle 2/3 full, not potting soil) • Sod patch (about 10 cm x 30 cm) • 3 plastic cups (about 16 oz.) • Leaves and twigs • 2 wire hangers • Blocks to support bottles • Sprinkling-style watering can • Tape • 1L water
- Carefully cut off one side of each bottle. Twist wire hangers around the necks of the bottles to connect them.
- Fill each bottle halfway with soil. Pat the soil down. Leave the soil in one bottle bare. Add twigs and leaves to cover the soil in one bottle to simulate forest soil cover. Cover the soil in another bottle with the sod patch.
- Suspend the bottles at a 25-degree angle with the spouts facing downward and over the cups.
- Use the watering can to slowly sprinkle equal amounts of water (about 330 mL each) evenly over the surface in each bottle.
• Describe the erosion in the bottles. • If each cup was a lake, in which would you choose to swim? • Which would be best for fish? • How would compressing the soil before adding water change the result?
Supplemental Worksheet :
www.soils4teachers.org/esw
Optional: Quantify the soil lost from each bottle.
- Measure the mass of the empty collection cups (no water or soil).
- After collecting runoff water, allow the water to evaporate completely from each cup.
- Measure the mass of each cup again.
- Subtract the mass of the empty cups step 1) from the mass of cups with dried soil (step 3). The difference is equal to the mass of soil eroded from each bottle.
- Science & Environment
- History & Culture
- Opinion & Analysis
- Destinations
- Activity Central
- Creature Features
- Earth Heroes
- Survival Guides
- Travel with AG
- Travel Articles
- About the Australian Geographic Society
- AG Society News
- Sponsorship
- Fundraising
- Australian Geographic Society Expeditions
- Sponsorship news
- Our Country Immersive Experience
- AG Nature Photographer of the Year
- View Archive
- Web Stories
- Adventure Instagram
Home Education Resources Science Experiment – Erosion
Science Experiment – Erosion
About this experiment
Forces of nature like wind, rain, and ice can all cause soil erosion. In this experiment, you’ll see how these natural patterns weather your soil landscape in different ways.
- 3 plastic containers (about the size of a shoe box is best)
- Soil (enough to fill your containers)
- A few ice cubes
- Spray bottle of water
- Optional: small stones, grass shoots or small plants
- In each container, pour soil up against one side, creating a sloped pile.
- In the first container, simulate water erosion by spraying water over the soil. What happens to the water? What happens to the soil?
- In the next container, simulate wind erosion by blowing air through the straw onto the top of the hill. How does the soil move, and how much of it moves? Do you think you would have had the same results on soil on flat ground, or on soil surrounded by trees?
- In the last container, simulate glacial erosion by placing a few ice cubes in the top of the hill, then pushing them down the hill. How is the landscape affected this time?
- Overall, examine which type of erosion had the greatest effect on the original structure of the soil. You could repeat the steps of this experiment, adding other materials to your “environments” like small plants or stones. How might the erosion differ in a more complex landscape?
Do you want to keep learning? Find more experiments here !
School with the Australian Reptile Park – Potoroo
Everything you need to know about our potoroos.
School with the Australian Reptile Park – Emus
School with the australian reptile park – elvis, watch latest web stories.
Birds of Stewart Island / Rakiura
Endangered fairy-wrens survive Kimberley floods
Australia’s sleepiest species
2024 Calendars & Diaries - OUT NOW
Our much loved calendars and diaries are now available for 2024. Adorn your walls with beautiful artworks year round. Order today.
In stock now: Hansa Soft Toys and Puppets
From cuddly companions to realistic native Australian wildlife, the range also includes puppets that move and feel like real animals.
Subscribe to our free Newsletter! →
Home › News
How electrified beaches could save our shores from devastating erosion
By StudyFinds Staff
Reviewed by Steve Fink
Research led by Alessandro Rotta Loria, Northwestern University
Aug 24, 2024
Beach erosion after storm activity along Australia's Gold Coast (© DavidJMorgan - stock.adobe.com)
CHICAGO — As coastal communities around the world grapple with the growing threat of erosion, researchers have uncovered a surprising potential solution: electricity. Scientists from Northwestern University say that applying mild electrical currents to marine sands could create natural, sustainable defenses against the relentless assault of waves and rising sea levels.
Coastal erosion is a pressing global issue, with approximately 40% of the world’s population living in coastal areas. As sea levels rise and storms intensify, nearly 26% of the Earth’s beaches could vanish by the end of this century. Traditional methods to combat this problem, such as seawalls and beach replenishment, often provide only temporary relief and can be costly to maintain.
Now, Northwestern’s Alessandro Rotta Loria and his team have found a way to turn sand into a rock-like substance, creating a natural armor for our coasts that could last for generations.
“My aim was to develop an approach capable of changing the status quo in coastal protection — one that didn’t require the construction of protection structures and could cement marine substrates without using actual cement,” Rotta Loria says in a statement. “By applying a mild electric stimulation to marine soils, we systematically and mechanistically proved that it is possible to cement them by turning naturally dissolved minerals in seawater into solid mineral binders — a natural cement.”
This innovative technique, called electrodeposition , draws inspiration from nature’s own architects: clams, mussels, and other shell-dwelling sea creatures . Just as these marine mollusks use dissolved minerals in seawater to construct their protective homes, the researchers harnessed these same minerals to create a natural cement between grains of sand. The twist? Instead of relying on the slow process of metabolic energy, they used electricity to kickstart the reaction.
In their laboratory experiments, the team applied mild electrical currents (ranging from 2 to 4 volts) to seawater-saturated sand. The results, published in Communications Earth & Environment , were nothing short of miraculous. The electricity triggered chemical reactions that converted naturally occurring ions and minerals in the seawater into solid calcium carbonate—the same stuff seashells are made of—and other mineral compounds like magnesium hydroxide and hydromagnesite.
These newly formed minerals act like a super-glue , binding sand particles together into a solid, immovable mass. “After being treated, the sand looks like a rock,” Rotta Loria explains. “It is still and solid, instead of granular and incohesive. The minerals themselves are much stronger than concrete, so the resulting sand could become as strong and solid as a sea wall.”
The researchers conducted a series of laboratory experiments using custom-designed electrochemical cells filled with silica sand and artificial seawater. By varying the voltage applied, the density of the sand, and the duration of the electrical treatment, they were able to observe how different conditions affected the formation and distribution of mineral deposits within the sand.
The results were striking. Depending on the voltage applied, the researchers observed the formation of different types of minerals. At lower voltages, calcium carbonate (the same material found in seashells and coral reefs ) was the primary deposit. As the voltage increased, magnesium hydroxide became more prevalent. Surprisingly, the team also discovered the formation of hydromagnesite, a mineral not previously observed in similar electrodeposition studies.
These mineral deposits didn’t just coat the sand particles; they formed bridges between them, effectively creating a cohesive, rock-like material from what was once loose sand. The strength of these newly formed “rocks” varied based on the amount of mineral content, with some samples achieving strengths of several megapascals – comparable to weak concrete.
Perhaps most importantly, the electrodeposition process significantly reduced the sand’s permeability. In some cases, the hydraulic conductivity (a measure of how easily water can flow through the material) decreased by an order of magnitude. This could have profound implications for mitigating erosion, as it would make the treated sand much more resistant to water infiltration and the undermining effects of waves and currents.
The implications of this discovery are vast. Traditional methods of coastal protection, such as building sea walls or injecting cement into the ground, are not only expensive but often temporary solutions. Sea walls themselves can erode, and the sand beneath them can liquefy, causing these multi-million dollar structures to sink into the very ground they’re meant to protect. In contrast, the electrodeposition method works with nature, strengthening the existing beach structure from within.
What’s more, this process is remarkably cost-effective. The researchers estimate that electrically cementing a cubic meter of sand would cost between $3 to $6, compared to up to $70 for conventional methods using binders. It’s also environmentally friendly and reversible—if needed, the process can be undone by simply switching the electrodes.
But perhaps the most exciting aspect of this research is its potential for widespread application. Beyond strengthening beaches, this technique could be used to stabilize sand dunes, reinforce seabeds beneath existing structures, and even heal cracks in reinforced concrete infrastructure along the shore.
However, as with any new technology, there are challenges to overcome before this method can be widely implemented. The researchers note that the process is sensitive to variables such as sand density and ionic concentration in the seawater. Scaling up from laboratory conditions to real-world beaches will require further study and refinement.
Despite these hurdles, the promise of this approach is clear. As climate change continues to threaten coastal areas with rising sea levels and more frequent storms, innovative solutions like electrodeposition could play a crucial role in protecting our shores. By harnessing the power of electricity and the minerals naturally present in seawater, we may have found a way to work with nature, rather than against it, in our ongoing battle against coastal erosion.
Paper Summary
Methodology.
The researchers conducted their experiments using custom-designed glass cells filled with various types of sand (including silica, calcareous, and iron sands) and artificial seawater. They applied different voltages (2.0, 3.0, and 4.0 volts) to the sand using electrodes, varying the duration of the electrical treatment. Throughout the experiments, they monitored pH levels and electrical current. After treatment, the researchers used various analytical techniques, including scanning electron microscopy, X-ray diffraction, and thermogravimetric analysis, to examine the type, amount, and distribution of minerals formed in the sand. They also conducted tests to measure the strength and permeability of the treated sand.
The study found that applying electrical current to seawater-saturated sand resulted in the formation of mineral deposits that cemented sand particles together. Lower voltages (2.0-3.0V) primarily produced calcium carbonate, while higher voltages (4.0V) favored magnesium hydroxide and hydromagnesite formation. The treated sand showed significantly increased strength, with some samples becoming as solid as weak concrete. The process also greatly reduced the sand’s permeability, making it more resistant to water infiltration. The volume of sand affected by the treatment increased with voltage and duration, with longer treatments affecting larger areas.
Limitations
While promising, this study has several limitations. The experiments were conducted under controlled laboratory conditions, which may not fully replicate the complex environment of real beaches. The longest duration tested was 28 days, which may not be sufficient to understand long-term effects and stability of the mineral formations. The study used artificial seawater, which might not represent all coastal environments. Additionally, while the researchers predict the treatment should last for decades, long-term durability in real-world conditions has not yet been tested.
Discussion and Takeaways
This research presents a novel approach to coastal erosion mitigation that could be more sustainable, cost-effective, and adaptable than traditional methods. The ability to selectively produce different minerals by controlling voltage offers flexibility in tailoring the treatment to specific coastal environments. The reduction in sand permeability and increase in strength suggest that this method could significantly enhance beach resistance to erosion. The reversibility of the process and its potential eco-friendliness are additional advantages. However, the researchers emphasize that more work is needed to understand how this process would work in real-world conditions and at larger scales.
Funding and Disclosures
The study was funded by several organizations, including the Army Research Office and the Center for Engineering Sustainability and Resilience at Northwestern University. The researchers used facilities at Northwestern University’s NUANCE Center, IMSERC Crystallography facility, and the Jerome B. Cohen X-Ray Diffraction Facility, which receive support from various sources including the National Science Foundation. The authors declared no competing interests related to the study.
About StudyFinds Staff
StudyFinds sets out to find new research that speaks to mass audiences — without all the scientific jargon. The stories we publish are digestible, summarized versions of research that are intended to inform the reader as well as stir civil, educated debate. StudyFinds Staff articles are AI assisted, but always thoroughly reviewed and edited by a Study Finds staff member. Read our AI Policy for more information.
Our Editorial Process
StudyFinds publishes digestible, agenda-free, transparent research summaries that are intended to inform the reader as well as stir civil, educated debate. We do not agree nor disagree with any of the studies we post, rather, we encourage our readers to debate the veracity of the findings themselves. All articles published on StudyFinds are vetted by our editors prior to publication and include links back to the source or corresponding journal article, if possible.
Our Editorial Team
Editor-in-Chief
Chris Melore
Sophia Naughton
Associate Editor
Related Content
Longevity breakthrough: Scientists uncover key gene that extends lifespan
September 4, 2024
Nicotine pouches surging in popularity — Because users can’t recognize what they are
Where to Find the Best Healthcare in the World
Leave a reply cancel reply.
©2024 Study Finds. All rights reserved. Privacy Policy • Disclosure Policy • Do Not Sell My Personal Information
- Skip to main content
- Skip to primary sidebar
Second Grade teachers! Join me Around the 2nd Grade Kampfire on Facebook! JOIN HERE
- Facebook Group
- Search this website
Around the Kampfire
Elementary Teaching Blog
Last updated by Linda Kamp on May 25, 2024 • 3 Comments
Water Erosion Science Experiment: Save the Lighthouse!
This water erosion experiment for kids is a fun science activity that is easy to do in any classroom. Ideal for teaching second grade students about slow earth changes, all of the materials you will need are from Walmart and the dollar store. The PowerPoint lessons and student lab sheet I use are part of this landforms & earth changes science unit .
By doing this experiment students gain:
- lab experience creating a simulation
- recording observations
- collecting data
- carrying out an investigation to answer a scientific question
What is Erosion?
Erosion is the process of wearing away rocks and soil by water, wind, ice, and gravity. The erosion process is part of the rock cycle and forms some pretty interesting landforms like mountain peaks, valleys, and coastlines.
I use this erosion science experiment, that is really more of a simulation, for students to answer the question, “How do ocean waves affect a coastline?”.
Base the Lab on Natural Phenomena
Whenever possible I like to center science lessons around natural events. Give students background for saving their lighthouses by showing this video of a 120 year old lighthouse endangered by erosion. The 1 minute video shows how the lighthouse was moved to a safer location.
Introduce the lab by posing the question: “How does erosion endanger a lighthouse?”
PowerPoint lesson & lab
Water Erosion Science Experiment
- paper lighthouse template
- plastic paint pan liner or shallow plastic bin
- student lab sheet
- blue food coloring (optional)
Place students in small groups or partners with the above materials needed to carry out their experiment. Guide students through the following procedure.
1. Fill the top two-thirds of a paint pan liner with sand. The sand we used is for hermit crabs from the pet department at Walmart, however regular sand is ideal.
2. Press the sand to form a “shoreline”. The sand should be about 2 or 3 inches deep.
3. Gently pour water into the pan filling the bottom half.
Students draw a diagram of their model on the lab sheet telling what each material used represents. Observe the sand as the water is poured. Ask students, “What change do you see as water is poured into the pan?”. (The sand is absorbing the water).
4. Using a plastic spoon, gently move the spoon up and down in the water to create small waves.
5. Have students pause after a few minutes to draw the changes on their lab sheet.
Erosion lab sheet
Ask students to note what happens when the waves wash up against the sand. (The “shoreline” changes because water washes back into the pan, carrying sand with it, and depositing it in a new location.)
6. Continue making waves until the shoreline is eroded up to the lighthouse.
7. Lastly, students discuss ways they can help prevent or slow down the erosion of their coastline. (A possible solution is to add rocks to the edge of the shoreline.)
More Erosion Experiments and Lesson Plans
This simple erosion experiment is a fun way for students to see the causes and effects of coastal erosion when learning about slow earth changes. It is part of a complete NGSS Landforms & Earth Changes unit for 2nd grade that is also available in a similar digital version .
Click here to see the yearlong 2nd grade science series.
Be sure to pin this lab for later so you have it when you plan!
Visit this post for more landforms and earth changes science activities and experiments.
10 Hands-On Activities for Teaching Earth Changes & Landforms
Share this:
Back to School
Free Grammar Review Activities!
Start the year in grammar off right! Grab a full week of lesson plans, teaching slides, and printables to review and practice parts of speech!
You May Also Enjoy These Posts
Reader Interactions
March 20 at 7:48 pm
Hi Linda! This lesson is amazing. Is it possible to just purchase the lighthouse lesson? It’s perfect for my homeschool curriculum but I don’t need the whole bundle. Thank you!
March 28 at 8:05 am
Hi Micaela! I’m so glad you like the lesson! Unfortunately I don’t have the labs available separately.
Leave a Comment Cancel reply
Your email address will not be published. Required fields are marked *
Notify me of new posts by email.
Hello Friends
I’m Linda Kamp, a 20 year primary grade teacher with a passion for creating educational materials that excite students and make learning fun! I'm so glad you're here!
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: 22 August 2024
Electrodeposition of calcareous cement from seawater in marine silica sands
- Andony Landivar Macias 1 ,
- Steven D. Jacobsen ORCID: orcid.org/0000-0002-9746-958X 2 &
- Alessandro F. Rotta Loria ORCID: orcid.org/0000-0001-6584-7526 1
Communications Earth & Environment volume 5 , Article number: 442 ( 2024 ) Cite this article
4057 Accesses
356 Altmetric
Metrics details
- Climate-change mitigation
- Geochemistry
- Natural hazards
The erosion of marine sediments is a pressing issue for coastal areas worldwide. Established methods to mitigate coastal erosion fail to provide lasting and sustainable solutions to protect marine ecosystems. Here we demonstrate the application of mild electrical stimulations to precipitate calcareous mineral binders from seawater in the pores of marine soils via electrodeposition, an alternative approach to mitigating coastal erosion. Results of electrochemical laboratory experiments unveil that the polymorphs, precipitation sites, intrusion mechanisms, and effects of electrodeposited minerals in marine sands vary as a function of the magnitude and duration of applied voltage, soil relative density, and electrolyte ionic concentration. Surprisingly, in addition to the precipitation of calcium carbonate and magnesium hydroxide, the formation of hydromagnesite is also observed due to electrically driven fluctuations in the local \({pH}\) . These electrodeposits lead to enhanced mechanical and hydraulic properties of the marine sands, indicating that electrodeposition routes could be developed to reinforce marine soils in coastal areas that more closely mimic natural systems.
Similar content being viewed by others
Mineralogy, morphology, and reaction kinetics of ureolytic bio-cementation in the presence of seawater ions and varying soil materials
Direct currents stimulate carbonate mineralization for soil improvement under various chemical conditions
A study of the effects of early diagenesis on the geotechnical properties of carbonate sediments (North West Shelf, Australia)
Introduction.
Coastal areas support the world’s most heavily developed regions 1 and house approximately 40% of the global population 2 . Yet, coastal areas face daunting challenges in the wake of extreme weather and rising sea-level, particularly erosion 1 , 3 , 4 .
Conventional methods to mitigate coastal erosion include seawalls 4 , 5 and beach replenishment 6 , 7 . However, these approaches are effective only within timeframes of a few years, representing costly and temporary solutions that require recursive retrofit 8 . In contrast, natural systems, such as coral reefs 7 , 9 and ecological barriers 7 , 10 , 11 , not only enhance coastal resilience, but also provide more structural support and can even contribute to beach accretion 8 .
Inspired by the way marine organisms utilize metabolic energy to grow their skeletons and shells through mineral precipitations in seawater, capable of resisting even the most extreme perturbations, this work explores an emerging and potentially disruptive approach to mitigate the erosion of marine soils. This approach consists of using electrical energy to precipitate comparable mineral precipitations to those that build the skeletons and shells of marine organisms in the pores of marine soils for cementation purposes and the ultimate enhancement of the erosion resistance of such materials.
The possibility to precipitate solid mineral binders in seawater through the application of mild electrical stimulations derives from the large buffering capacity 12 , 13 and wide availability of ions 14 , 15 characterizing such an electrolyte, and the process of electrodeposition 16 : the electrically mediated precipitation of minerals dissolved in solutions. Specifically, when an electrical current is applied to seawater, a variety of reduction and oxidation (redox) reactions occur, along with solid precipitations 17 . These reactions commence with the release of hydroxide ions ( \(O{H}^{-}\) ) around cathodic interfaces, leading to an increase in the local \({pH}\) . Under these conditions, the generated hydroxide ions react with naturally dissolved divalent cations in seawater ( \(M{g}^{2+}\) and \(C{a}^{2+}\) ) and bicarbonate anions ( \({HC}{O}_{3}^{-}\) ), yielding the otherwise non-spontaneous precipitation of two ubiquitous minerals: magnesium hydroxide ( \({Mg}{\left({OH}\right)}_{2}\) ) and calcium carbonate ( \({CaC}{O}_{3}\) ) 18 , 19 .
Currently, the electrodeposition of minerals in seawater is largely exploited to protect marine structures against corrosion 20 , 21 , 22 , contribute to marine life 16 , 23 , 24 , and heal cracks in shoreside infrastructure 25 , 26 . Recent experimental evidence shows that electrodeposition can also cement marine substrates in contact with metallic structures 27 and the minerals formed in this manner appear to be durable 28 . However, the understanding of the influence of electrodeposition on the structure and properties of marine soils remains limited. In this context, a knowledge gap exists due to the apparent absence of studies examining the influence of the magnitude of the applied voltage on mineral selectivity, composition, spatial distribution, intrusion mechanisms, and effects on soil properties, despite voltage being the driver of electrodeposition. Although a mechanistic understanding of the reactions and products of electrodeposition has been achieved for seawater 17 , this knowledge is unavailable for soils, where the electrical and kinetic phenomena ruling electrodeposition are inherently more complex due to the influence of geometric, physical, and chemical constraints exerted by the pore network of such materials. Therefore, the implications of variable reaction regimes and mineral formations that can be achieved with different electrochemical potentials on the structure and properties of marine soils remain uncharted.
The fundamental hypothesis of this work is that the electrodeposition in marine soils depends principally on the features of the (1) electrical treatment, (2) soil fabric, and (3) electrolyte solution. To test this overarching hypothesis, we carried out two distinct types of experiments in custom-designed electrochemical cells on silica sand under highly controlled conditions (see Methods). The first involved short-term batch experiments with no water recirculation to uncover the selectivity of electrodeposits in soils. The second comprised long-term experiments with periodic water recirculation to unveil the intrusion mechanisms and effects of electrodeposits in soils under conditions comparable to open seawater environments. Altogether, these experiments explored the influence of the following central variables for the features of applied electrical treatments, treated porous materials, and electrolyte solutions: (1) magnitude and duration of the applied voltage, (2) soil relative density, and (3) electrolyte ionic concentration, respectively. Multiple physical, chemical, and mechanical characterization methods were used to investigate the morphology, structure, composition, spatial distribution, and effects of electrodeposited minerals in sand.
The results of this work uncover key mechanisms and effects of the artificial precipitation of minerals in marine soils, advancing geochemical knowledge and supporting engineering and technology in the development of more effective mitigation strategies for coastal erosion.
Physical characterization
Physical characterization achieved via scanning electron microscopy (SEM), Raman spectroscopy, and imaging analyses of the electrodeposited minerals highlights key microscopic features of their precipitation loci, morphology, and polymorphism, along with apparent macroscopic effects on the silica sand (Fig. 1 , Methods). The first uncovered pattern is that the precipitation loci and type of electrodeposits depend on the applied voltage, which governs the reaction regimes and hence the electrodeposition process (Fig. 1a ). At a relatively low voltage (2.0 V), sparse electrodeposits surround sand particles without significantly bonding them together (Fig. 1a , row 1, column 1). At a medium voltage (3.0 V) and high voltage (4.0 V), denser electrodeposits surround and bond sand particles via mineral bridges (Fig. 1a , row 1, columns 2 and 3). The SEM results consistently indicate two dominant mineral formations: \({CaC}{O}_{3}\) in the forms of calcite (rhombohedral morphology) and aragonite (needle-like morphology) 29 , as well as \({Mg}{\left({OH}\right)}_{2}\) in the form of brucite (fibrous compounds constituted of lamellar structures) 18 , 19 . \({CaC}{O}_{3}\) formations appear especially at lower voltages (Fig. 1a , rows 2 and 3, columns 1 and 2), whereas \({Mg}{\left({OH}\right)}_{2}\) formations are mainly electrodeposited at higher voltages (Fig. 1a , rows 2 and 3, column 3); when both \({CaC}{O}_{3}\) and \({Mg}{\left({OH}\right)}_{2}\) are electrodeposited, \({Mg}{\left({OH}\right)}_{2}\) minerals coat the silica soil particles and form a substrate for the precipitation of \({CaC}{O}_{3}\) minerals (Fig. 1a , rows 2 and 3).
a SEM images at different scales (i.e., particle, interface, and mineral scale) of loose silica sand subjected to short-term electrical conditioning (i.e., 7 days) at voltages of 2.0, 3.0, and 4.0 V; b SEM images of all the visible solid minerals achieved by electrodeposition; c Raman spectra for minerals precipitated in the tested sands, compared with reference spectra; d SEM images on silica sand at different relative densities (i.e., loose and dense) at a voltage of 2.0 V; e Summary of the affected volumes of cemented (or affected) sand at voltages of 2.0, 3.0, and 4.0 V for short-term tests (7 days) and long-term tests (28 days); f Photographs of the effects of the distinct intrusion mechanisms associated with variable voltages for the long-term tests at 2.0, 3.0, and 4.0 V.
An analysis of SEM images of selected samples of the tested silica sand shows that \({Mg}{\left({OH}\right)}_{2}\) is formed as both lamellar and block brucite, and \({CaC}{O}_{3}\) is found in the increasingly stable forms of vaterite, aragonite, and calcite (with aragonite dominating in quantity with respect to calcite and vaterite) (Fig. 1b ). Notably, a quantitative analysis of Raman spectra of precipitated minerals on the silica sands consistently indicates \({CaC}{O}_{3}\) precipitations in the forms of calcite and aragonite, as well as \({Mg}{\left({OH}\right)}_{2}\) in the form of brucite; additionally, Raman spectra also unveil the precipitation of hydromagnesite, \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) (Fig. 1c and Supplementary Fig. 1 ). XRD analyses reported in the sequel (Fig. 2e ) also quantitatively confirm such evidence.
A second discernible pattern in the electrodeposition of minerals in marine sands is that the size and precipitation loci of electrodeposits depend on the relative density (Fig. 1d ), which is a proxy of porosity. Loose sands exhibit smaller minerals and more uniform and widespread precipitations (Fig. 1d , columns 1 and 2). In contrast, dense sands exhibit larger minerals and less uniform and widespread precipitations, as well as more aragonite compared with calcite, with almost no brucite (Fig. 1d , columns 3 and 4).
An analysis of the volume and mass of “cemented” (or affected) sand elucidates a third discernible feature for the electrodeposition in marine soils (Fig. 1e ); the volume of soils influenced by electrodeposits depends on the duration and magnitude of the applied electrical stimulation. After 7 days of treatment, a limited volume of sand is cemented by electrodeposits. The largest cemented volume is achieved for 3.0 V, whereas smaller volumes are achieved (in order) for 4.0 and 2.0 V. After 28 days, a larger volume of sand is cemented by electrodeposits. The largest cemented volume is achieved for 4.0 V, whereas smaller volumes are achieved (in order) for 3.0 and 2.0 V. For a given voltage, the affected volume of electrodeposits decreases with an increase in the soil relative density, whereas it increases for a longer electrical stimulation. Extending the duration of the electrical stimulation from 7 to 28 days does not significantly change the extent of the cemented volume of sand for 2.0 and 3.0 V, whereas it tremendously enhances such extent for 4.0 V. After 7 days, the soil area affected by electrodeposition reaches a radial distance of about 2–3 times the electrode diameter for 3.0 V. In contrast, this area exceeds a radial distance of 20 times the electrode diameter after 28 days of treatment with 4.0 V.
Consideration of the macroscopic effects of electrodeposition for varying magnitudes of voltage elucidates a final feature for the electrodeposition of minerals in marine soils (Fig. 1f ); there exist distinct intrusion mechanisms and extents of volumes affected by electrodeposition, mainly as a function of changes in the electrochemical reactions ruling electrodeposition (see Discussion and Supplementary Discussion). The application of a low voltage (2.0 V) over a prolonged time yields electrodeposits near the cathode, which penetrate in the soil farther away from the electrode for a medium voltage (3.0 V) and influence a significant portion of the bulk of the soil for a high voltage (4.0 V). Comparable results are obtained within shorter timeframes (Supplementary Fig. 2 ), with the difference that the overall volume influenced by electrodeposits is consistently less significant. As electrodeposits grow within sands, they form a cemented material. Electrical current and resistance vary according to such mineral formations, with an overall decrease of the former and an increase of the latter due to the constant applied potential over time (Supplementary Fig. 3 ).
Chemical characterization
Chemical characterization of the environmental conditions in the sand and the electrodeposits (Fig. 2 ) achieved by \({pH}\) measurements, energy dispersive X-ray spectroscopy (EDS), X-ray diffraction (XRD), and thermogravimetric analyses (TGA) (see Methods) supports an enhanced analysis of the mineral properties and their formation in closed and open electrochemical systems. The temporal variations of \({pH}\) at the anode (Fig. 2a ) and cathode (Fig. 2b ) measured in the short-term in a closed system show that the \({pH}\) in the sand can change significantly, leading to the creation of an acidic environment near the anode and a more alkaline environment near the cathode due the influence of oxidation and reduction reactions, respectively. For both loose and dense sands, the low voltage level (i.e., 2.0 V) results in minimal to negligible changes in \({pH}\) . However, for the highest voltage level (i.e., 4.0 V) the impact is considerable. In all cases, \({pH}\) variations are delayed in dense sands due to their smaller porosity and greater tortuosity compared to loose sands. Upon the termination of the electrical conditioning, the \({pH}\) starts to return to its initial value. In contrast, the temporal variations of \({pH}\) at the anode and cathode (Fig. 2c ) measured in the long-term tests in an open system highlight minimal to negligible variations in \({pH}\) due to the periodic recirculation of seawater. By refilling the system with \(C{a}^{2+}\) and \(M{g}^{2+}\) ions, these conditions keep the bulk electrolyte ionic concentration approximately constant.
a Evolution of \({pH}\) over time near the anode for short-term experiments (7 days), measured via a \({pH}\) meter; b Evolution of \({pH}\) over time near the cathode for short-term experiments (7 days), measured via a \({pH}\) meter; c Evolution of \({pH}\) over time in the electrochemical cell and at the inlet and outlet of the cell for long-term experiments (28 days), measured via a \({pH}\) meter; d Summary of the elemental ratio of calcium over magnesium \(({Ca}/{Mg})\) obtained for loose and dense sands at different voltages (2.0, 3.0, and 4.0 V) for both short-term and long-term tests through EDS mapping (error bars indicate standard deviations); e Summary of the relative weight of mineral according to Rietveld refinement deriving from XRD analyses of sands ( \({m}_{c}\) is the mass cement content, with \({M}_{c}\) the mass of the cementing calcareous deposits and \({M}_{i}\) the mass of the cemented soil; error bars indicate standard deviations); f Summary of the relative weight of minerals according to TGA measurements for sands subjected to long-term electrical conditioning (error bars indicate standard deviations).
An analysis of the elemental mapping obtained via EDS for loose and dense sands (Fig. 2d and Supplementary Fig. 4 ) reveals a decrease in the \({Ca}/{Mg}\) ratio with increasing voltage, consistently with all the qualitative results obtained via SEM (Fig. 1a ). The same trend for the \({Ca}/{Mg}\) ratio as a function of the applied voltage is calculated from the Rietveld refinement of the XRD spectra (Fig. 2e and Supplementary Fig. 5 ) and the TGA results (Fig. 2f and Supplementary Fig. 6 ). The quantitative differences between the obtained trends for the \({Ca}/{Mg}\) ratio are associated with the different techniques employed to analyze this variable, remaining qualitatively consistent. Notably, both the XRD and TGA results confirm again the dominant precipitation of calcium-based minerals at limited voltages, as opposed to the dominant precipitation of magnesium-based minerals at substantial voltages, as previously shown by the SEM analyses (Fig. 1a ). The XRD analyses substantiate the presence of vaterite, calcite, aragonite, brucite, and hydromagnesite, as previously highlighted by the Raman spectra (Fig. 1c and Supplementary Fig. 1 ).
Hydromechanical characterization
The nucleation and growth of mineral electrodeposits change the structure of marine sands. Consequently, these mineral precipitations modify the properties of such materials. The results of this work quantitatively unveil changes in the hydraulic conductivity, \(k\) , and the unconfined compressive strength, \({UCS}\) , of marine sands subjected to electrodeposition (Fig. 3 ).
a Variation in hydraulic conductivity as a function of applied voltage level for both loose and dense sands subjected to long-term electrical conditioning; b Variation in unconfined compressive strength against mass cement content for both loose and dense sand (the samples tested refer to long-term electrical conditioning tests involving the application of 3.0 and 4.0 V; the trendline refers to variations in unconfined compressive strength achieved by microbially induced calcite precipitation in a myriad of coarse-grained soils, including sands 38 ). Error bars indicate standard deviations.
A hydraulic characterization of the sands subjected to long-term electrical conditioning (Fig. 3a ) indicates a decreasing trend in hydraulic conductivity for materials subjected to increasing voltages. Interestingly, the decrease in hydraulic conductivity of a cemented sand compared to a clean (i.e., untreated) sand can be as high as one order of magnitude for the highest voltage (4.0 V). In all cases, the results show a higher hydraulic conductivity of loose compared to dense sands because of the more porous packing of the former compared to the latter.
An analysis of the correlation between the unconfined compressive strength and the mass cement content of the tested sands (Fig. 3b ) shows that both loose and dense sands benefit from a marked increase in strength with the mass cement content. The electrodeposits can yield cemented sands with \({UCS}\) reaching several MPa. In other words, electrodeposition can turn initially cohesionless sands into rocks.
Mechanistic analysis of electrodeposition reactions and effects
The results unveil a selectivity of the mineral type, polymorph, proportions, precipitation loci, and intrusion mechanisms achieved by electrodeposition in marine sands depending on the (1) applied voltage, (2) soil relative density, and (3) electrolyte ionic concentration.
The applied voltage fundamentally influences the process and effects of electrodeposition in sands by triggering different electrochemical reactions that govern the kinetics of mineral precipitations and their penetration in the bulk of the treated materials. An analysis of the applied voltages with respect to a reference electrode (see Supplementary Discussion, Supplementary Figs. 7 and 8 , and Supplementary Table 1 ) indicates that two key redox reactions govern the electrodeposition in sands: the oxygen reduction reaction (ORR) and the water reduction reaction (WRR). The identification of the reaction regimes corresponding to the application of 2.0, 3.0, and 4.0 V specifically allows identifying the reaction pathways that result in the electrodeposition of \({{Mg}\left({OH}\right)}_{2}\) , \({{CaCO}}_{3}\) , and \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) in the experiments.
At the lower voltage of 2.0 V, which is equivalent to −0.84 \({V}_{{Ag}/{AgCl}}\) or −0.64 \({V}_{{SHE}}\) , ORR involves \({{OH}}^{-}\) production as follows:
At the higher voltages of 3.0 and 4.0 V, which correspond to −1.32 \({V}_{{Ag}/{AgCl}}\) or −1.13 \({V}_{{SHE}}\) as well as −1.58 \({V}_{{Ag}/{AgCl}}\) or −1.38 \({V}_{{SHE}}\) , respectively, WRR shifts selectivity of \({{OH}}^{-}\) production through hydrogen evolution, thereby leading to the formation of hydrogen gas bubbles at the cathode surface. Under neutral and alkaline conditions, hydrogen evolution proceeds as follows:
Under acidic solutions, hydrogen evolution proceeds instead as follows:
The direct consequence of the generation of \({{OH}}^{-}\) ions in the vicinity of a cathode, whether in either the ORR or WRR regimes, is an increase in the interfacial \({pH}\) 17 . Due to this phenomenon, naturally occurring \(M{g}^{2+}\) ions in seawater will react with the released \({{OH}}^{-}\) ions, leading to the precipitation of insoluble \({{Mg}\left({OH}\right)}_{2}\) as follows:
In other words, upon the application of electrical stimulations to sands wetted by seawater under ambient conditions, \({{Mg}\left({OH}\right)}_{2}\) will precipitate initially instead of \({{CaCO}}_{3}\) due to the higher concentration of \({{Mg}}^{2+}\) ions compared to \({{Ca}}^{2+}\) ions in seawater. Under these conditions, the mineral formation process is specifically driven by kinetics instead of thermodynamics. Following a decrease in the local \({pH}\) and concentration of \({{Mg}}^{2+}\) ions in solution resulting from the precipitation of \({{Mg}\left({OH}\right)}_{2}\) , two possible outcomes will subsequently occur.
On the one hand, in the ORR regime, \({{CaCO}}_{3}\) will precipitate atop the initial \({{Mg}\left({OH}\right)}_{2}\) formations due to a favorable local \({pH}\) that is nonetheless insufficient for continued \({{Mg}\left({OH}\right)}_{2}\) precipitations. In this context, \({{Ca}}^{2+}\) and \({{HCO}}_{3}^{-}\) ions naturally present in seawater will react with the released \({{OH}}^{-}\) ions, resulting in the precipitation of insoluble \({{CaCO}}_{3}\) as follows:
Alternatively, \({{CaCO}}_{3}\) precipitation will also possibly occur due to the influence of the released \({{OH}}^{-}\) ions on the carbonate equilibrium:
which will lead to a higher concentration of carbonate ions ( \({{CO}}_{3}^{2-}\) )
and the consequent precipitation of insoluble \({{CaCO}}_{3}\) as follows:
On the other hand, in the WRR regime, \({{Mg}\left({OH}\right)}_{2}\) will continue to precipitate instead of \({{CaCO}}_{3}\) due to a sufficiently high local \({pH}\) driven by hydrogen evolution. Consistent with results presented here for soils wetted by seawater and studies of electrodeposition in seawater 21 , 27 , 30 , the \({Ca}/{Mg}\) ratio will hence decrease with increasing values of the applied voltage. However, as the \({pH}\) of seawater is typically of 8.2-8.4 31 under standard conditions but the \({pH}\) of stability of brucite is of 9.4 under ambient conditions 32 , brucite will turn into aragonite over time 28 . Notably, in the WRR regime, the penetration of electrodeposits in the pore network of soils will be facilitated by the formation a hydrogen gas front that pushes the electrodeposits away from the cathode.
In addition to \({{Mg}\left({OH}\right)}_{2}\) and \({{CaCO}}_{3}\) , the results of this work surprisingly support the electrodeposition of \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) in soils wetted by seawater. To the best of our knowledge, the identification of hydromagnesite appears unprecedented in the literature about electrodeposition in seawater 17 , 18 , as only \({{Mg}\left({OH}\right)}_{2}\) and \({{CaCO}}_{3}\) have been traditionally identified. Interestingly, the presence of dissolved silica in the soil-water system considered here may play an important for this evidence, as silica is renowned to play a catalytic role in the development of magnesium bearing carbonates 33 . \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) preferentially forms in the ORR regime and in the early stages of WRR regime, characterized by relatively slow kinetics. This evidence, quantitatively supported by the results of Raman spectroscopy and XRD analyses, is attributed to successive \({pH}\) and salinity fluctuations associated with the alternating formation of \({{Mg}\left({OH}\right)}_{2}\) and \({{CaCO}}_{3}\) observed in this regime, as opposed to the sustained \({pH}\) leading to the preferential growth of only \({{Mg}\left({OH}\right)}_{2}\) , which is characteristic of the fully established WRR regime due to hydrogen evolution. Under these conditions, \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) can form due to the local availability of \(M{g}^{2+}\) , \(C{O}_{3}^{-}\) , and \(O{H}^{-}\) ions in solution, which results in the precipitation of such an hydrated form of magnesium carbonate, \({MgC}{O}_{3}\) . The precipitation of \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) , instead of \({MgC}{O}_{3}\) , takes place at ambient temperature because it is kinetically favored due to the high hydration character of \(M{g}^{2+}\) ions in solution 34 . However, hydromagnesite is a metastable mineral form that will eventually transform (through dehydration and recrystallization) into anhydrous magnesite, either at elevated temperatures or under high \(C{O}_{2}\) pressures 35 . This phenomenon is anticipated to further enhance the strength of electrodeposited sands, as \({MgC}{O}_{3}\) has better mechanical properties than \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) .
Complementary oxidation reactions proceed at the anode, which balance the reduction reactions of water at the cathode. In the ORR, oxygen evolution leads to:
In the WRR regime, oxygen evolution develops in alkaline media as follows:
whereas in acidic media as follows:
Higher applied voltage leads to higher production rates of \({H}_{2}\) and \({{OH}}^{-}\) at the cathode and \({O}_{2}\) and \({H}^{+}\) at the anode, increasing the driving force for \({H}^{+}\) and \({{OH}}^{-}\) ions to diffuse further away from the electrodes at which they are produced. This phenomenon can result in a crosstalk of ions between the electrodes and neutralization of the \({{OH}}^{-}\) intended to participate in mineral precipitation as follows:
In addition to the applied voltage, the soil relative density also influences polymorph selectivity because changes in the relative density of soils affect the available surface area and number of nucleation sites, which are renowned to govern mineral formations 36 , 37 . A looser packing provides a constrained number of ample nucleation surfaces and a facilitated frequency of ionic interactions, which lead to smaller and uniformly distributed electrodeposits. In contrast, a denser packing provides more numerous and smaller nucleation surfaces and a hampered frequency of ionic interactions, which involve larger and less uniform electrodeposits. Increasingly tortuous and denser soils, coupled with the influence of low voltages, specifically appear to involve a slow reaction kinetics that results in alternating \({{Mg}\left({OH}\right)}_{2}\) and \({{CaCO}}_{3}\) formations (with a smaller amount of precipitated \({{Mg}\left({OH}\right)}_{2}\) compared to looser soils subjected to the same voltages).
The electrolyte ionic concentration also contributes to polymorph selectivity. The \({pH}\) variations observed in the closed batch experiments are significant and involve a progressive ionic deprivation of seawater that affects the mineral precipitations. In contrast, minimal \({pH}\) variations and substantial mineral precipitations are observed in the long-term experiments where seawater is periodically recirculated, and the ionic concentrations of calcium and magnesium ions remain approximately constant. This evidence is considered representative of open seawater conditions due to the continuous replenishment of ions through marine currents and biogenic activity.
Implications for coastal erosion mitigation
The results show that electrodeposition changes the structure of marine sands. As the porous structure of sands progressively fills with newly formed electrodeposits, the porosity is reduced while particle contacts increase and are cohesively bonded. As a result, the hydraulic conductivity of marine soils decreases whereas their shearing strength increases.
The variations in hydraulic conductivity and strength achieved with electrodeposition are consistent in magnitude with those of other emerging methods of soil cementation 38 , 39 . Compared to such methods, treatments using electrodeposition benefit nonetheless from their avoidance of external fluid injections and their ability to trigger mineral precipitations in porous networks of highly variable sizes. These advantages result from the pervasive action of the electrically mediated charge transport and mineral precipitation pathways leveraged by electrodeposition. Additionally, treatments targeting the electrodeposition of minerals via electricity deriving from renewable energy sources offer the benefit of producing green hydrogen in the WRR regime, which may be collected for valuable purposes. In principle, a downside of electrodeposition in seawater is that it will lead to chlorine production at significant potentials in the WRR regime 17 , 40 , which may affect some fish and marine mammals, especially in closed environments 24 . In practice, the chemistry of seawater favors dominant oxygen production over chlorine, and any chlorine gas will rapidly neutralize with dissolved organic matter in any open environment 24 . Therefore, treatments using electrodeposition have significant potential to engineer marine soils sustainably.
In summary, the ability to tailor the hydraulic conductivity and strength of marine sands with electrodeposition appears to bear tremendous relevance for coastal erosion mitigation, as these material properties significantly influence erosion patterns, sediment transports and seepage, and the structural stability of coastal and offshore structures.
Conclusions
Motivated by the lack of sustainable and lasting approaches to mitigate coastal erosion worldwide, this study presented an experimental laboratory study on the electrically mediated precipitation of mineral binders in marine sands via the process of electrodeposition. Specifically, by harnessing custom-designed electrochemical cells, this study systematically and mechanistically explored the electrodeposition of calcium- and magnesium-based minerals in silica sand saturated by seawater. For the first time, this work analyzed the influence of the (1) magnitude and duration of the applied voltage, (2) soil relative density, and (3) electrolyte ionic concentration, which are central variables for any applied electrical treatment, treated porous material, and electrolyte solution, respectively.
The results of this work shed light on distinct reaction mechanisms governing the electrodeposition of \({CaC}{O}_{3}\) , \({Mg}{\left({OH}\right)}_{2}\) , and \(M{g}_{5}{\left(C{O}_{3}\right)}_{4} {\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) in sands saturated by seawater. A transition from oxygen reduction to water reduction particularly occurs between 2.0 and 3.0 V, with marked hydrogen evolution at 4.0 V. During oxygen reduction, hydroxide formation is mass transport limited and results in predominant \({CaC}{O}_{3}\) depositions, mostly in the form of calcite, with companion \(M{g}_{5}{\left(C{O}_{3}\right)}_{4}{\left({OH}\right)}_{2}\cdot 4{H}_{2}O\) formations. During water reduction, hydroxide formation is reaction rate limited and favors \({Mg}{\left({OH}\right)}_{2}\) productions, mostly in the form of brucite. By identifying these two distinct potential regimes, this work highlights that control of the applied electrode potential, with consideration of the soil relative density and electrolyte features, enables selectivity of the type of precipitated mineral and polymorph, mineral precipitation location, and mineral intrusion mechanism. Therefore, electrodeposition allows achieving variable effects on the structure and properties of electrodeposited sands. These effects are limited in the oxygen reduction reaction regime. In contrast, they are remarkable in the water reduction reaction regime, where hydrogen evolution enables higher mineral production rates by preventing surface film buildup and electrode passivation, as well as the creation of a gas front that pushes electrodeposits in the soil bulk, affecting significant volumes of soils around cathodic interfaces.
This work advances the knowledge of mineral electrodeposition in soils saturated by seawater, underscoring the feasibility of harnessing such a process to protect coastal areas against erosion processes. Additionally, the results of this work enhance the analysis of problems related to sedimentation, biomineralization, and geological carbon sequestration, which are governed by the nucleation and growth of minerals in porous materials.
Experimental laboratory tests
This study performed an integrated experimental investigation across different spatial scales (e.g., from the macroscale of sand specimens to the microscale of sand particles) and temporal scales (i.e., from 7 to 28 days of electrical conditioning). Specifically, this work performed two classes of electrochemical laboratory experiments in custom-designed electrochemical cells, followed by several qualitative and quantitative characterization methods to address the physical and chemical properties of the electrodeposits, their spatial distribution and features in the porous structure of the treated soils, and their resulting effects on the hydraulic conductivity and shearing strength of such materials.
Short-term batch experiments (lasting 7 days) with no water recirculation were performed to assess the fundamental mechanisms of electrodeposition in soils. Long-term recirculation experiments (lasting 28 days) with periodic water recirculation were performed to simulate electrodeposition under field conditions and explore the interconnected influence of structural changes on the bulk properties of soils. Each short-term and long-term experiment was repeated three times to ensure repeatability. Among the various features that characterize any (1) applied electrical treatment, (2) treated porous material, and (3) electrolyte solution, these experiments allowed to address the following variables that were hypothesized to play a central role for the electrodeposition of minerals in marine sands:
Magnitude of applied voltage. This variable was chosen due to its prominent influence on the electrodeposition of minerals in seawater 18 , which was unchartered for soils saturated by seawater at the time this work was performed. Voltage levels of \(E=\) 2.0, 3.0, and 4.0 V were selected to address the efficacy and mineral precipitation mechanisms at relatively low, moderate, and high potentials, respectively. These voltage values also correspond to \(E=-0.839,-1.323,\) and \(-1.579\) \({V}_{{Ag}/{AgCl}}\) , which correspond to development of the sole oxygen reduction reaction, the onset of the water reduction reaction, and the dominant development of the water reduction reaction, respectively (see Supplementary Discussion).
Duration of applied voltage. This variable was chosen to explore the spatial growth of electrodeposits over time and assess the presence of a “curing time” for such deposits, which is renowned to characterize other methods for soil cementation resorting to carbonate precipitations 39 . Test durations of \(t=\) 7 and 28 days were chosen for this purpose.
Soil relative density. This variable was chosen due to its prominent influence on soil mechanics 41 and the efficacy of other treatments for soil cementation 39 through its impact on the contact topology and packing of granular materials, although this evidence was unchartered for soils subjected to electrodeposition at the time this work was performed. Relative densities associated with loose states ( \({D}_{r}=\) \(25\pm 5\) \(\%\) ) and dense states ( \({D}_{r}=\) \(80\pm 5\) \(\%\) ) were selected for this purpose. Analyzing the soil relative density also allowed to indirectly assess the influence of surface density, which is a critical parameter for the formation of minerals 36 . Surface density refers to the concentration or availability of sites that are suitable for nucleation, where mineral depositions can prosper 41 . Higher surface densities generally translate to a greater number of active sites available for deposition, while lower surface densities indicate a fewer number of available active sites for deposition 37 .
Electrolyte ionic concentration. This variable was again chosen based on its prominent influence on the electrodeposition of minerals in seawater 18 , although this evidence was unchartered for soils subjected to electrodeposition at the time this work was performed. This variable was indirectly assessed by developing short- and long-term electrical conditioning experiments without or with a periodic replenishment of seawater, respectively. This approach ensured a variable (i.e., decreasing) and quasi-constant (i.e., stable) ionic concentration of seawater during electrodeposition, given that its underlying reactions consume key ions over time if no mass transfer with a source of fresh seawater takes place.
The experimental protocol underpinning the developed electrochemical experiments was as follows. Initially, the sand was compacted in the relevant electrochemical cell to achieve the desired relative density through the pluviation method 42 . Next, the compacted sand was slowly saturated with seawater, by injecting such a fluid at minimal velocity from hose connector located at the bottom of the electrochemical cells to prevent the formation of air bubbles and prevent any noteworthy soil disturbance, which would have otherwise affected the relative density. Then, the sand was allowed to rest for 1 h while ensuring that it achieved full saturation. Afterward, voltage was applied to the electrochemical cell while keeping drainage lines (i.e., the hoses) closed. After the electrochemical conditioning, the lid atop the system was removed and the bottom hose was open to allow for desaturation for 72 h to facilitate the subsequent sample extraction process. Once extracted, the samples were thoroughly washed with deionized water to eliminate any excess sodium chloride ( \({NaCl}\) ) and eventually dried for characterization or testing. The same experimental procedure characterized the short- and long-term experiments, at the exception that seawater was recirculated (i.e., flushed out from the top hose) using the peristaltic pump over a period of 8 h to ensure approximately constant \({pH}\) conditions during each long-term experiment. In the long-term experiments, \({pH}\) measurements were continuously taken not only inside the cell as in the short-term experiments, but also at the inlet and outlet of seawater flow.
Tested materials
The tested material in this study was F35 sand—a coarse, rounded silica sand composed of over 99% SiO 2 quarried at Ottawa, Illinois purchased from U.S. Silica. Supplementary Fig. 9 shows the particle size distribution obtained via a sieve analysis as well as optical microscopy and scanning electron microscopy (SEM) images to illustrate qualitative features of the tested material. Supplementary Table 2 summarizes further details of the properties of the tested material. Silica sand was chosen due to its chemical inertness with respect to acid, bases, oxidation, and reduction owing to the strong silicon-oxygen bond present in its crystal structure 43 .
The tested sand was saturated with artificial seawater to create an electrolyte solution as comparable as possible to actual seawater 14 . Artificial seawater was prepared by mixing distilled Type 1 ultrapure water ( \(18.2\) \({\mbox{M}}\omega -{\mbox{cm}}\) of resistivity at 25 °C from Millipore Sigma) with a “sea salt” preparation mixture (from Lake Products Company LLC) according to the ASTM D1141-98 44 . This process consisted in the dissolution of 41.953 g of sea salt mixture with enough water to make 1 L of total solution. The salts, once mixed and dissolved, yielded a fluid with a chemical composition of over 99% similarity with natural seawater 44 . The \({pH}\) of the prepared solution was measured and adjusted as necessary with weak bases or acids to ensure similarity with natural conditions (i.e. \(,\) \({pH}=\) 8.2–8.4 31 ).
Experimental apparatuses
For this study, two classes of custom-designed tempered glass electrochemical cells of different volume capacities were employed (Supplementary Fig. 10 ): (1) 2-L capacity cells and (2) 500-mL capacity cells. The 2-L liter cell (Supplementary Fig. 10a ) is cylindrical with an inner diameter \(D=160\) \({\mbox{mm}}\) , an inner height of \(H=120\) \({\mbox{mm}}\) , a thickness \(t=5\) \({\mbox{mm}}\) , a flanged top of \(15\) \({\mbox{mm}}\) , and two hose connectors with an opening of 9 \({\mbox{mm}}\) located at heights of \({H}_{1}=20\) \({\mbox{mm}}\) and \({H}_{2}=70\) \({\mbox{mm}}\) . Similarly, the 500 mL cell (Supplementary Fig. 10b ) is cylindrical with an inner diameter \(D=85\) \({\mbox{mm}}\) , an inner height of \(H=100\) \({\mbox{mm}}\) , a thickness \(t=5\) \({\mbox{mm}}\) , a flanged top of \(15\) \({\mbox{mm}}\) , and two serrated hose connectors with an opening of 9 \({\mbox{mm}}\) located at heights of \({H}_{1}=20\) \({\mbox{mm}}\) and \({H}_{2}=70\) \({\mbox{mm}}\) . Depending on the experiments, the hose connectors in the cells were used to saturate the tested material with the chosen electrolyte solution and/or to run hydraulic conductivity measurements with an Intlab peristaltic pump. This pump provided a precise liquid dosing of 30 ml/min within the equipment admissible flowrates of 5–50 mL/min, with an accuracy of \(\pm 1\) mL/min.
Alongside the developed glass cells, Teflon lids with O-rings were used to seal the cells and ensure a closed environment limiting evaporation and chemical reactions with the atmosphere. The lids of both cells incorporated holes to host the anode, the cathode, two \({pH}\) meters, and one gas outlet, ensuring optimal monitoring of the environmental conditions in the cells. The cell and lid materials were chosen given their low electrical conductivity and low chemical reactivity to prevent any external contamination into the testing chamber.
The electrode materials were selected to ensure ideal conditions for studying electrodeposition. For the anode material, a platinum rod (with a diameter \(D_{a}=0.78{{{\rm{mm}}}}\) , a total length of \({L}_{a}\,={\mbox{}}100{\mbox{mm}}\) , and an embedded length of \({L}_{a}^{* }=\) 70 and 45 mm in the batch and recirculation experiments, respectively) was chosen to prevent oxidation and ensure a highly controlled flow of electricity. For the cathode material, a steel 316 rod ( \({D}_{c}=2.0\) \({\mbox{mm}}\) , \({L}_{c}={\mbox{}}100{\mbox{mm}}\) , and \({L}_{c}^{* }=\) 70 mm for the batch tests; \({D}_{c}=3.0\) \({\mbox{mm}}\) , \({L}_{c}={\mbox{}}100{\mbox{mm}}\) , and \({L}_{c}^{* }=\) 45 mm for the recirculation tests) was chosen for its widespread use in practice. Different electrode embedment lengths were used in the batch and recirculation tests to ensure equivalent current density. All electrochemical measurements gathered in the core electrochemical experiments of this work used a Tektronix Keithley Series 2280 S High Precision Bech Power supply that allowed the measurement of voltage and current. Companion electrochemical measurements gathered in complementary electrochemical experiments employed a Biologic SP 150e potentiostat equipped with electrochemical impedance spectroscopy and the setup detailed elsewhere 17 .
Two \({pH}\) meters (Hanna Instruments soils \({pH}\) meter HI2002) with an accuracy of \({pH}=\pm 0.01\) were utilized to measure \({pH}\) at 1.5 cm from each electrode. A calibration of these instruments was performed with buffer solutions of \({pH}=\) 4.01, 7.01, and 10.01 from Hanna Instruments.
Supplementary Figs. 10c and 10d shows the experimental set-up for the short-term and long-term electrical conditioning experiments, respectively.
Physical, chemical, hydraulic, and mechanical characterization
Each physical, chemical, hydraulic, and mechanical characterization of the results was repeated at least three times to ensure representativeness and reliability of the data. Errors bars are presented wherever relevant and applicable to show the variability in the obtained results.
Scanning electron microscopy (SEM) was employed to qualitatively study the microscopic features of the minerals and their spatial distribution in the porous network of the sand (Fig. 1a, b, d ). SEM analyses employed a Quanta 650 F microscope at an accelerating voltage of 20 kV. Before subjecting to a scanning electron microscope, the electrodeposits were coated with gold ( \({Au}\) ) in a 20 nm thickness using a Denton’s Desk IV deposition system. The results encompass different scales: the particle scale, showcasing whole sand particles; the interface scale, revealing the contacts between sand and mineral particles; and the mineral scale, focusing on the electrodeposits.
Energy dispersive X-ray spectroscopy (EDS) was employed to qualitatively assess the elemental composition of the materials and create elemental maps (Supplementary Fig. 4 ). This technique facilitated the identification of electrodeposited minerals and the detection of possible impurities caused by oxidation reactions of the electrodes. Yet, EDS elucidated the analysis of the spatial distribution of elements, thereby providing valuable insights into the microscopic behavior of the material and the mineral intrusion mechanisms. Although it is acknowledged that EDS cannot provide data representative of the three-dimensionality of the precipitation patterns in the tested sands, repeatability analyses confirmed the high consistency in the obtained data, which were also qualitatively in agreement with X-ray diffraction (XRD) and thermogravimetric analyses data (TGA). As XRD and TGA analyses were performed on randomly selected mineral samples from the tested soils and inherently provide representative data of the three-dimensionality of the precipitations, the consistency between all these results provides strong evidence of the soundness of the developed analyses and interpretations.
Raman spectroscopy was carried out using a custom-built, confocal micro-Raman spectrometer with a 300 mW, 458-nm excitation laser (Melles Griot 85-BLS-601). The system was built around an Olympus-BX optical microscope, 0.3-meter spectrograph (Andor Shamrock 303i), and Newton DU970 EMCCD camera from Andor Technology. Neutral density filters were added to reduce the laser power to ~10 mW at the sample. The excitation laser was focused to ~1–2 μm spot size through a 100x objective (Mitutoyo M Plan Apo SL100x) with 13.0 mm working distance and a 0.55 numerical aperture. Spectra were collected from 0–4000 cm −1 Raman shift with an acquisition time of 3 seconds, averaged over 10 acquisitions. The Raman spectra are shown alongside reference spectra at low and high wavenumber regions, enabling a precise identification of calcium- and magnesium-based minerals (Fig. 1c ). Supplementary Fig. 1 identifies the characteristic peaks of the studied substances for detailed reference, representing a quantitative and complementary piece of evidence in support of the analysis of the obtained results.
Image processing analyses were used to examine the macroscopic features of the electrodeposited sand, with a focus on the volume affected by electrodeposits (Fig. 1e, f , and Supplementary Fig. 2 ). The volume calculations were performed analytically through the open-source code Fiji ImageJ 45 . The volume calculations were addressed in three key steps. First, any original image was digitalized. Second, such an image was binarized into two colors (i.e., black and white) to better classify the area of the analyzed view. Such area ( \({A}_{s}\) ) was calculated by the software, and subsequently it was converted into an equivalent cylindrical shape. For this process, the height of the equivalent cylinder was assumed to be the length of the electrode (i.e., \({L}_{c}=100\) \({\mbox{mm}}\) ), and the equivalent diameter, \({D}_{{eq}}\) , was calculated as follows:
Finally, based on the geometrization of the irregular volume of electrodeposited sand into an equivalent cylinder, the cemented soil volume ( \({V}_{e}\) ) was calculated without taking in account the volume of the used electrode.
XRD analyses were employed to quantitatively assess the composition of the crystal structure and phase composition of electrodeposited minerals, allowing for an accurate detection of polymorphs (Fig. 2e and Supplementary Fig. 5 ). XRD was performed on samples of electrodeposited sand collected near the cathode for all combinations of voltage levels, relative densities, and test durations. XRD analyses on samples indicated a mineral content ranging from 8.6% up to 96.5%. Powder XRD patterns were collected at room temperature on a STOE-STADI-P powder diffractometer equipped with an asymmetrically curved Germanium monochromator (CuKα1 radiation, λ = 1.54056 Å) and one-dimensional silicon strip detector (MYTHEN2 1 K from DECTRIS) to identify the crystalline phases of the deposits. The line-focused Cu X-ray tube was operated at 40 kV and 40 mA. The fine powder deposits were packed in an 8 mm metallic mask and sandwiched between two polyimide layers of tape. Intensity data from diffraction angles of 15° to 75° (2θ) were collected for 1 h per sample. The instrument was calibrated against a NIST Silicon standard (640d) before the measurement. Phase identification was made by Rietveld refinements using the Profex analysis software 46 . All the reference spectra were used from the Crystallography Open Database (COD). Specifically, COD 9016706, COD9000229, COD 9015898, COD 9002348, and COD 9007620 spectra were used, corresponding to calcite, aragonite, vaterite, brucite, and hydromagnesite, respectively. The spectrum for quartz was also gathered from The American Mineralogist Crystal Structure Database (AMCSD) and plotted for reference with the label AMCSD 0012866.
TGA was performed to quantitatively assess the mass and composition of electrodeposited minerals (Supplementary Fig. 6 ). For these experiments, the samples were first grinded into a fine powder (i.e., \( < 4\) \(\mu {{{\rm{m}}}}\) ) using a mortar and pestle and eventually subjected to the analyses. These analyses used a Netzsch STA 449 F3 Jupiter Simultaneous Thermal Analysis instrument purging nitrogen at 50 mL/min. 50 mg of the electrodeposits were placed in a 0.35 ml crucible made of \(A{l}_{2}{O}_{3}\) with a weight of 200 mg. TGA results were calculated based on the mass loss at different stages of heating. A heating rate of 10 °C/min was used over a temperature range of 40–1000 °C. The weight loss between 200–500 °C was associated with the destruction of \(O{H}^{-}\) bonds, which are associated with magnesium hydroxide. The weight loss between 500–900 °C was associated with the release of \(C{O}_{2}\) , which is associated with calcium carbonate. The weights at 200 °C ( \({M}_{200}\) ) and 500 °C ( \({M}_{500}\) ) were determined from the TGA curves. The difference between the weights in the 200–500 °C range ( \({\Delta M}_{200-500^{\circ}{{{\rm{C}}}}}={M}_{500}-{M}_{200}\) ) gave the weight of the generated water gas. A conversion factor between the atomic weight of magnesium hydroxide (58.32 g/mol) and that of water (18.02 g/mol) needs to be applied and leads to the calculation of mass of magnesium hydroxide as follows:
Similarly, the weights at 500 °C ( \({M}_{500}\) ) and 900 °C ( \({M}_{900}\) ) were determined from the TGA curves. The difference between the weights in the 500–900 °C range ( \({\Delta M}_{500-900^{\circ}{{{\rm{C}}}}}={M}_{900}-{M}_{500}\) ) gave the weight of the generated carbon dioxide gas. A conversion factor between the atomic weight of carbon dioxide (44.01 g/mol) and that of calcium carbonate (100.09 g/mol) needs to be applied and leads to the calculation of mass of calcium carbonate as follows:
The evaluation of mineral mass content was performed through acid washing tests 47 . These tests involved a \(5\pm 0.1\) g of sample mixed with 20 mL of a 1 M solution of hydrochloric acid ( \({HCl}\) ) to dissolve electrodeposits. The remaining solution and insoluble solid was washed with Type 1 ultrapure distilled water on a Fisherbrand P4 grade filter paper for 10 min. The mass cement content was calculated as follows:
where \({M}_{w}\) is the mass of the sample post acid wash, \({M}_{i}\) is the original mass of the electrodeposited sand sample, and \({M}_{c}\) is the mass of the cementing calcareous electrodeposits. The TGA curves allow for the development of derivative thermogravimetric (DTG) curves, which represent the first derivative of the TGA curve. The DTG curves are particularly valuable to determine inflection points and key variations in weight for deeper interpretations during analyses.
Constant head permeability tests were performed to measure the hydraulic conductivity of the tested sand before and after electrical conditioning (Fig. 3a ). The experiments devoted to assessing the hydraulic conductivity of the tested loose and dense sands not subjected to any electrical conditioning were performed right after the preparation of such materials in the cell, without using such materials for any other purpose due to the possible structural disturbance of the sands subjected to a hydraulic gradient for these tests. The experiments devoted to assessing the hydraulic conductivity of the tested loose and dense sands were performed after the application of the electrical conditioning. In this case, reference was made to the long-term experiments because they yielded materials benefitting from a more substantial cementation, which facilitated the quantification of the considered parameter. These tests followed the procedure prescribed by ASTM D3424-19 48 .
Unconfined compressive strength tests were performed to quantify the strength of electrodeposited sands (Fig. 3b ). While other tests, such as flume tests, could be considered more relevant to characterize the erosion resistance of materials, the unconfined compressive strength provides direct and first-order information about the erosion resistance of marine sediments, especially when samples of limited size are available 49 , 50 , 51 . These tests were performed on cubic sand samples extracted from the cell after the long-term tests (again to cope at best with the available volume of cemented materials). The tested samples were carefully cut after extraction from the testing cell (i.e., clearly, these sand samples did not include the electrodes inside them but were collected around the electrodes) and had dimensions of 10 mm inside to comply the requirements of a Representative Elementary Volume 52 . In alignment with the requirements provided by ASTM D7012-04 53 , the dimension of these samples was determined in a way to be sufficiently larger than the average particle size of the cemented soils, thereby providing reliable results. Prior to testing, the electrodeposited sand samples were laid on a gypsum capping to avoid stress concentrations and interconnected ill-conditioning of the results. Unconfined compressive strength tests followed the procedure prescribed by ASTM D7012-04 53 and were performed with a MTS model C43.104 mechanical testing apparatus.
Data availability
The quantitative data shown in the figures are publicly available at Zenodo: https://doi.org/10.5281/zenodo.12741685 .
Luijendijk, A. et al. The state of the world’s beaches. Sci. Rep. 8 , 6641 (2018).
Article Google Scholar
Mentaschi, L., Vousdoukas, M. I., Pekel, J.-F., Voukouvalas, E. & Feyen, L. Global long-term observations of coastal erosion and accretion. Sci. Rep. 8 , 12876 (2018).
Zhang, K., Douglas, B. C. & Leatherman, S. P. Global warming and coastal erosion. Clim. Change 64 , 41–58 (2004).
van Rijn, L. C. Coastal erosion and control. Ocean Coast. Manag. 54 , 867–887 (2011).
Masria, A., Iskander, M. & Negm, A. Coastal protection measures, case study (Mediterranean zone, Egypt). J. Coast. Conserv. 19 , 281–294 (2015).
Dean, R. G. Beach Nourishment: Theory and Practice . vol. 18 (World Scientific, 2002).
Gracia, A., Rangel-Buitrago, N., Oakley, J. A. & Williams, A. T. Use of ecosystems in coastal erosion management. Ocean Coast. Manag. 156 , 277–289 (2018).
Goreau, T. J. & Trench, R. K. Innovative Methods of Marine Ecosystem Restoration . (CRC Press, 2012).
Fernando, H. J. S., Samarawickrama, S. P., Balasubramanian, S., Hettiarachchi, S. S. L. & Voropayev, S. Effects of porous barriers such as coral reefs on coastal wave propagation. J. Hydro-Environ. Res. 1 , 187–194 (2008).
Barbier, E. B. Natural barriers to natural disasters: replanting mangroves after the tsunami. Front. Ecol. Environ. 4 , 124–131 (2006).
Borsje, B. W. et al. How ecological engineering can serve in coastal protection. Ecol. Eng. 37 , 113–122 (2011).
Frankignoulle, M. A complete set of buffer factors for acid/base CO 2 system in seawater. J. Mar. Syst. 5 , 111–118 (1994).
Jiang, L.-Q., Carter, B. R., Feely, R. A., Lauvset, S. K. & Olsen, A. Surface ocean pH and buffer capacity: past, present and future. Sci. Rep. 9 , 18624 (2019).
Article CAS Google Scholar
Kester, D. R., Duedall, I. W., Connors, D. N. & Pytkowicz, R. M. Preparation of artificial seawater. Limnol. Oceanogr. 12 , 176–179 (1967).
Anthoni, J. F. The chemical composition of seawater. Magnesium 2701 , e9062 (2006).
Google Scholar
Hilbertz, W. Electrodeposition of minerals in sea water: experiments and applications. IEEE J. Ocean. Eng. 4 , 94–113 (1979).
Devi, N. et al. Mechanistic insights into electrodeposition in seawater at variable electrochemical potentials. Adv. Sustain. Syst. 8 , 2300446 (2024).
Carré, C. et al. Electrochemical calcareous deposition in seawater. A review. Environ. Chem. Lett. 18 , 1193–1208 (2020).
Hartt, W. H., Culberson, C. H. & Smith, S. W. Calcareous deposits on metal surfaces in seawater—a critical review. Corrosion 40 , 609–618 (1984).
Neville, A. & Morizot, A. P. Calcareous scales formed by cathodic protection—an assessment of characteristics and kinetics. J. Cryst. Growth 243 , 490–502 (2002).
Yao, C. et al. Formation of calcareous deposits in the tidal zone and its effect on cathodic protection. Npj Mater. Degrad. 7 , 30 (2023).
Yang, Y., Scantlebury, J. D. & Koroleva, E. V. A study of calcareous deposits on cathodically protected mild steel in artificial seawater. Metals 5 , 439–456 (2015).
Goreau, T. J. F. & Prong, P. Biorock electric reefs grow back severely eroded beaches in months. J. Mar. Sci. Eng. 5 , 48 (2017).
Goreau, T. J. Marine electrolysis for building materials and environmental restoration. Electrolysis , Ch. 13 (eds Linkov, V. & Kleperis, J) 273–290 (IntechOpen, 2012).
Abe, M., Fukute, T., Yokota, M. & Sasaki, H. Applicability of electrodeposition as a repairing method for deteriorated marine concrete structures. Rep. Port Harb. Res. Inst. 30 , 41 (1991).
Rotta Loria, A. F., Shirole, D., Volpatti, G., Guerini, A. & Zampini, D. Engineering concrete properties and behavior through electrodeposition: a review. J. Appl. Electrochem. 53 , 193–215 (2023).
Carré, C. et al. Electrochemical limestone synthesis in seawater binds metal grids and sediments for coastal protection. Environ. Chem. Lett. 18 , 1685–1692 (2020).
Zadi, L. et al. Physico-chemical stability evaluation of a sedimentary agglomerates use for the coastal protection. J. Coast. Conserv. 27 , 12 (2023).
Ševčík, R., Šašek, P. & Viani, A. Physical and nanomechanical properties of the synthetic anhydrous crystalline CaCO 3 polymorphs: vaterite, aragonite and calcite. J. Mater. Sci. 53 , 4022–4033 (2018).
Rousseau, C., Baraud, F., Leleyter, L., Jeannin, M. & Gil, O. Calcareous deposit formed under cathodic protection in the presence of natural marine sediments: A 12 month experiment. Corros. Sci. 52 , 2206–2218 (2010).
Marion, G. M. et al. pH of seawater. Mar. Chem. 126 , 89–96 (2011).
Pokrovsky, O. S. & Schott, J. Experimental study of brucite dissolution and precipitation in aqueous solutions: surface speciation and chemical affinity control. Geochim. Cosmochim. Acta 68 , 31–45 (2004).
Hobbs, F. W. C. & Xu, H. Magnesite formation through temperature and pH cycling as a proxy for lagoon and playa paleoenvironments. Geochim. Cosmochim. Acta 269 , 101–116 (2020).
Deelman, J. C. Breaking Ostwald’s rule. Chem. Erde-Geochem. 61 , 224–235 (2001).
CAS Google Scholar
Königsberger, E., Königsberger, L.-C. & Gamsjäger, H. Low-temperature thermodynamic model for the system Na2CO3 − MgCO3 − CaCO3 − H2O. Geochim. Cosmochim. Acta 63 , 3105–3119 (1999).
Godinho, J. R. A., Gerke, K. M., Stack, A. G. & Lee, P. D. The dynamic nature of crystal growth in pores. Sci. Rep. 6 , 33086 (2016).
Bang, J.-H. et al. Specific surface area and particle size of calcium carbonate precipitated by carbon dioxide microbubbles. Chem. Eng. J. 198–199 , 254–260 (2012).
Landivar Macias, A. & Rotta Loria, A. F. SANISAND-C*: Simple ANIsotropic constitutive model for SAND with Cementation. Int. J. Numer. Anal. Methods Geomech . 47 , 2815–2847 (2023).
Terzis, D. & Laloui, L. A decade of progress and turning points in the understanding of bio-improved soils: a review. Geomech. Energy Environ. 19 , 100116 (2019).
Dresp, S., Dionigi, F., Klingenhof, M. & Strasser, P. Direct electrolytic splitting of seawater: opportunities and challenges. ACS Energy Lett. 4 , 933–942 (2019).
Mitchell, J. K. & Soga, K. Fundamentals of Soil Behavior . (Wiley, New York, 2005).
Vaid, Y. P. & Negussey, D. Preparation of reconstituted sand specimens. Adv. Triaxial Test. Soil Rock ASTM STP 977 , 405–417 (1988).
Greenwood, N. N. & Earnshaw, A. 9 - Silicon. in Chemistry of the Elements (Second Edition) 328–366 (Butterworth-Heinemann, Oxford, 1997). https://doi.org/10.1016/B978-0-7506-3365-9.50015-8 .
American Society for Testing and Materials. ASTM D1141-98. Practice for the Preparation of Substitute Ocean Water . https://doi.org/10.1520/D1141-98R21 (1998).
Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9 , 676–682 (2012).
Doebelin, N. & Kleeberg, R. Profex: a graphical user interface for the Rietveld refinement program BGMN. J. Appl. Crystallogr. 48 , 1573–1580 (2015).
Choi, S.-G., Park, S.-S., Wu, S. & Chu, J. Methods for calcium carbonate content measurement of biocemented soils. J. Mater. Civ. Eng. 29 , 06017015 (2017).
American Society for Testing and Materials. ASTM D2434-19 Standard Test Method for Permeability of Granular Soils (Constant Head) . (ASTM international, 2019).
Thorne, C. R. Field measurements of rates of bank erosion and bank material strength. Eros. Sediment Transp. Meas. 133 , 503–512 (1981).
Kamphuis, J. W. & Hall, K. R. Cohesive material erosion by unidirectional current. J. Hydraul. Eng. 109 , 49–61 (1983).
Skafel, M. G. & Bishop, C. T. Flume experiments on the erosion of till shores by waves. Coast. Eng. 23 , 329–348 (1994).
Al-Raoush, R. & Papadopoulos, A. Representative elementary volume analysis of porous media using X-ray computed tomography. Powder Technol. 200 , 69–77 (2010).
American Society for Testing and Materials. ASTM. D7012-04 Standard Test Method for Compressive Strength and Elastic Moduli of Intact Rock Core Specimens under Varying States of Stress and Temperatures . (West Conshohocken, PA, USA., 2004).
Download references
Acknowledgements
The authors would like to thank Prof. Jean-François Gaillard, Dr. Nishu Devi, and Dr. Yeong-Man Kwon for their insightful comments on this work. Dr. Yeong-Man Kwon is particularly thanked for contributing to the XRD analyses performed as a part of this study. Dr. Raul Marrero is thanked for his help with the training on the sample preparation and mechanical testing procedures under unconfined conditions. This work made use of the EPIC facility of Northwestern University’s NUANCE Center, which has received support from the SHyNE Resource (NSF ECCS-2025633), the IIN, and Northwestern’s MRSEC program (NSF DMR-1720139), of the IMSERC Crystallography facility at Northwestern University, which has received support from the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF ECCS-2025633), and Northwestern University, and the Jerome B.Cohen X-Ray Diffraction Facility supported by the MRSEC program of the National Science Foundation (DMR-1720139) at the Materials Research Center of Northwestern University and the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF ECCS-1542205.). The seed funding provided by the Center for Engineering Sustainability and Resilience of Northwestern University in support of early developments behind this work is appreciated. This study is part of a broader research program funded by the Army Research Office (Grant No. W911NF2210291), whose financial support is greatly acknowledged.
Author information
Authors and affiliations.
Department of Civil and Environmental Engineering, Subsurface Opportunities and Innovations Laboratory, Northwestern University, 2145 Sheridan Road, Evanston, IL, 60208, USA
Andony Landivar Macias & Alessandro F. Rotta Loria
Department of Earth and Planetary Sciences, Northwestern University, 2145 Sheridan Road, Evanston, IL, 60208, USA
Steven D. Jacobsen
You can also search for this author in PubMed Google Scholar
Contributions
A.F.R.L. conceived the study idea and with A.L.M. designed the experimental apparatus and procedures. A.L.M. carried out the experiments and data acquisition during cementation. A.L.M., S.D.J., and A.F.R.L. conducted sample characterizations and data reduction. All authors contributed equally to interpreting the data and writing the manuscript.
Corresponding author
Correspondence to Alessandro F. Rotta Loria .
Ethics declarations
Competing interests.
The authors declare no competing interests.
Peer review
Peer review information.
Communications Earth & Environment thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editors: Olusegun Dada and Carolina Ortiz Guerrero. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Peer review file, supplementary information, rights and permissions.
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, 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 you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. 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-nc-nd/4.0/ .
Reprints and permissions
About this article
Cite this article.
Landivar Macias, A., Jacobsen, S.D. & Rotta Loria, A.F. Electrodeposition of calcareous cement from seawater in marine silica sands. Commun Earth Environ 5 , 442 (2024). https://doi.org/10.1038/s43247-024-01604-3
Download citation
Received : 04 May 2024
Accepted : 06 August 2024
Published : 22 August 2024
DOI : https://doi.org/10.1038/s43247-024-01604-3
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
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.
Stream Table Erosion Lab
I love teaching weathering and erosion! One of the best ways to learn about weathering and erosion is to SEE it in action. There are a few ways to do this:
- Have students go on a hunt outside and look for evidence of weathering or erosion. Have them take a picture, predict what caused the weathering or erosion, and upload to your LMS. (This is a great homework assignment if you don’t want to do it around campus).
- Have them explore cool landforms around the world that have been formed by weathering or erosion on Google maps. (If you want a premade lesson on this, check out this interactive diagram on TpT or on my website ).
- Have them play with stream tables!
Large stream tables are expensive, and a pain to store from year to year. You can easily make your own (for a lot less money!) by using small paint roller pans. I’ve done this lab multiple times over the years and worked out the kinks for you. Here are my best tips!
How to build your own stream table
(If you are more of a video kinda person and would like to SEE this lab instead of read about it, check out this youtube tutorial instead).
First, you’re going to need to head to Lowes or Home Depot. You have a few options when purchasing paint roller pans: there are metal and plastic. You’re going to want to purchase plastic paint roller pans that stand up at an angle themselves, not the really flexible insert pans. You can see in the picture below, the bottom has raised ridges that help it sit at an angle. They cost roughly $2 each, so I spent $16 on 8 of them (don’t worry, they will last you multiple years).
Next, you need to drill holes along the bottom of the tray for drainage. I drilled 5 per tray, and that was more than adequate. If you don’t want lakes to form at the bottom of the tray, drill the holes towards the bottom. If you are fine with water pooling, drill them about halfway up the pan.
The last step is to fill them with sand. If you have sand around your campus or neighborhood you can use for free, have at it! You can also purchase sand at Home Depot for about $5 a bag. One bag was enough for me to fill all 8 trays.
The day before the lab, you will need to also prep 2 more things:
- Get one plastic cup per group. Drill or poke a hole in the bottom of the cup. Students will be filling these with water and allowing the water to drain out onto their mini stream table.
- Get an ice cube tray or two, fill the bottom with some small gravel, and top off with water. Freeze overnight. These will be used to represent glaciers.
Stream table experiment TIPS
These stream tables are obviously much shorter than the typical ones you can order from science suppliers. That limits the amount of erosion that can take place. Here are some tips to help them work the best:
- Start with wet sand, and pack it down tight . If the sand is dry, it will all run down without creating a clear channel. It also helps to form a small, narrow channel with your hand before running any water down. Since the tables are so short, this helps the stream form faster.
- I’ve done this lab both indoors (when I had a lab space with lots of sinks and rubber hoses) and outside (when I had no sinks). I definitely recommend doing this outdoors to avoid messes and sand going down your drains. You will need a couple 5 gallon buckets- just fill them up and place outside. Multiple groups can share water from 1 bucket.
- Alright, lets talk water. If you simply pour water from a cup or beaker, the water tends to spread out too fast and channels don’t form. Solution: poke a hole in the bottom of a plastic cup . Have students cover the hole with their finger while they fill the cup, and then uncover the hole once they are ready for water to flow down the pan. A smaller, more direct water source helps the channel form more easily.
- Between trials, students need to dump out any excess water, flatten out the dirt, and re-form that shallow channel.
- In the past I’ve tried to have students blow through straws and see wind erosion, but it doesn’t work well with wet sand. If you would like to try this, I would have one tray that has very dry sand, and have groups rotate over to it.
Stream Table Variables
What can you have students test? What variables impact the rate of erosion?
- You can change the slope of the pan by holding it up at a steeper angle.
- You can have students change the water discharge by providing a second cup with a larger hole.
- You can have students place rocks or small plants down into the sand and see what impact it has on erosion. Does the water go around the rocks? Over the rocks?
- Have students gently push an ice cube glacier down the table. Does any sand below the glacier erode? What happens to the bottom of the ice cube by the time it reaches the bottom of the tray?
- Have students start their channel with meanders instead of a straight channel. Do meanders make the river erode faster or slower? On what part of the curve is erosion taking place?
While overall this is a very qualitative lab and there isn’t much data to record, it is fun for students to get outside and visualize stream formation. You can find a lab write-up on TpT or on my website .
- Read more about: Geology
Hi, I'm Becca!
Search the site, browse by category.
- A list of ALL blog posts
- Back to School
- Biochemistry
- Body Systems
- Classification
- Classroom Decor
- Classroom Management
- Distance Learning
- End of the School Year
- Experiments
- Field Trips
- For NEW Teachers
- Formative Assessment
- Media in the Classroom
- Microscopes
- Photosynthesis & Respiration
- Plate Tectonics
- Sustainability
- Teacher Tips
- Weather and Climate
Get Freebies!
You might also like....
Science Seek and Finds
Layers of the Earth Lessons
Teaching the Electromagnetic Spectrum
Want a fun way to practice science vocabulary? Try out seek and finds!
Privacy Overview
Cookie | Duration | Description |
---|---|---|
cookielawinfo-checkbox-analytics | 11 months | This cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Analytics". |
cookielawinfo-checkbox-functional | 11 months | The cookie is set by GDPR cookie consent to record the user consent for the cookies in the category "Functional". |
cookielawinfo-checkbox-necessary | 11 months | This cookie is set by GDPR Cookie Consent plugin. The cookies is used to store the user consent for the cookies in the category "Necessary". |
cookielawinfo-checkbox-others | 11 months | This cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Other. |
cookielawinfo-checkbox-performance | 11 months | This cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Performance". |
viewed_cookie_policy | 11 months | The cookie is set by the GDPR Cookie Consent plugin and is used to store whether or not user has consented to the use of cookies. It does not store any personal data. |
- Skip to main content
Join All-Access Reading…Doors Are Open! Click Here
- All-Access Login
- Freebie Library
- Search this website
Teaching with Jennifer Findley
Upper Elementary Teaching Blog
Erosion Science Experiment With Grass
It’s one thing to talk about scientific concepts and ideas in the classroom, but another to actually see them at work. If you’re looking for a way to give students a firsthand look at erosion, this is a great experiment to do so. In this science experiment, students will answer the question Does Grass Help Prevent the Erosion From Water?
Want to see more spring activities and resources for other subject areas?
Erosion Science Experiment Materials Needed:
- A foil baking pan
- Dirt with grass mixed in
- A pitcher or cup to pour water from
- Free passage and comprehension questions download (at the end of this post)
Erosion Science Experiment Directions
1. Place tightly packed clumps of sand, dirt, and grass/dirt into the foil pan.
2. Add water to a measuring cup. Make your predictions. Which pile – sand, dirt, or grass/dirt – will stand up to erosion the best?
3. Gently and slowly pour water over each of the piles.
4. You can tell from this photo that grass does help to prevent erosion from water. The clump of dirt with grass on top pretty much stayed in tack, with minor mudflow.
The Science Behind the Erosion Experiment
One element that contributes greatly to erosion is rain. Whether it’s a steady drizzle or a major downpour, rain can do a lot of damage. So, what helps slow the forces of soil erosion?
One of the things that binds together all different kinds of land is root systems. That’s why planted grasses are so helpful for erosion prevention. Grasses have soft stems and are better for binding than woody stems. They also have fibrous roots that spread out in all directions.
So, planted grass is excellent for erosion prevention because it helps bind the soil.
After the Experiment Reading Activity
I like to incorporate reading and writing into every science experiment, activity, or demonstration that we do and this is no exception.
For this activity, the students will read a short text that describes the science behind it (similar to what is explained above for the teacher’s reference). Then the students will use the details they learned in the text to explain what happened during the science experiment. They will also answer three comprehension questions using details from the text.
The questions your students will answer include:
- What are some things that contribute to erosion?
- How do fibrous roots help with soil erosion?
- If you wanted to keep soil in place, what are some things you could do to prevent soil erosion?
After reading the passage and answering the questions, you can invite your students to share their responses and have a classroom discussion about erosion.
How Can I Get the Free Printable?
Click here or on the image below to get the freebie that accompanies this erosion science experiment .
What are your favorite spring science activities or STEM activities? Let me know in the comments! I am always looking for new science experiments that my students will love.
If you want more resources and even freebies for spring , check out my recommended activities, such as math worksheets, egg hunt activities, review game ideas, printables for read alouds, engineering with jelly beans, science experiments involving flowers and seeds, and more.
Shop This Post
Science Centers with Reading Passages
Spring and Easter Paired Passages
Spring Reading Comprehension Passages
Share the knowledge, reader interactions, leave a comment cancel reply.
Your email address will not be published. Required fields are marked *
Notify me of follow-up comments by email.
Notify me of new posts by email.
You may also love these freebies!
Math Posters
Reading Posters
Morphology Posters
Grammar Posters
Welcome Friends!
I’m Jennifer Findley: a teacher, mother, and avid reader. I believe that with the right resources, mindset, and strategies, all students can achieve at high levels and learn to love learning. My goal is to provide resources and strategies to inspire you and help make this belief a reality for your students.
- News & Events
- Contact & Visit
- Faculty & Staff
- McCormick Advisory Council
- Departments & Institutes
- Diversity Data
- Faculty Journal Covers
- Areas of Study
- Bachelor's Degrees
- Music & Engineering
- Combined BS / MS Program Collapse Combined BS / MS Program Submenu
- Murphy Scholars Program Projects
- Undergraduate Honors
- Certificates & Minors Collapse Certificates & Minors Submenu
- Integrated Engineering Studies
- Engineering First® Program
- Theme Requirement
- Research Opportunities
- Personal & Career Development
- Global Opportunities
- Existing Groups
- McCormick Community
- Transfer AP/IB Credits
- ABET Course Partitioning
- Enrollment and Graduation Data
- Full-time Master's
- Part-time Master's
- MS with Interdepartmental Minors
- Application Checklist
- Application FAQs
- Financial Aid
- International Students
- Student Groups
- Career & Professional Development
- All Areas of Study
- Departments & Programs
- Apply to Northwestern Engineering
- Faculty Fellows
- Office of the Dean
- Administration, Finance, Facilities, & Planning
- Alumni Relations & Development
- Career Development
- Corporate Engagement
- Customer Service Center
- Faculty Affairs
- Global Initiatives
- Graduate Studies
- Information Technology
- Marketing & Communications
- McCormick Advising System
- Personal Development StudioLab
- Professional Education
- Research Offices
- Undergraduate Engineering
- Newsletter Signup
- Information for the Media
- Tech Room Finder
Fighting Coastal Erosion with Electricity
Bioinspired process makes marine sand more durable, resistant to erosion.
The Problem
With climate change and rising sea levels, coastal erosion is an increasing threat.
A new process that uses electricity to form a natural cement between grains of sand, transforming it into solid, immoveable rock.
Why It Matters
This strategy could offer a lasting, inexpensive, and sustainable solution for strengthening global coastlines.
Professor Alessandro Rotta Loria; Andony Landivar Macias, former PhD candidate in Rotta Loria’s laboratory; Steven Jacobsen, Weinberg College of Arts and Sciences professor
New research from Northwestern University has systematically proven that a mild zap of electricity can strengthen a marine coastline for generations — greatly reducing the threat of erosion in the face of climate change and rising sea levels.
In the new study, researchers took inspiration from clams, mussels and other shell-dwelling sea life, which use dissolved minerals in seawater to build their shells.
Similarly, the researchers leveraged the same naturally occurring, dissolved minerals to form a natural cement between sea-soaked grains of sand. But, instead of using metabolic energy like mollusks do, the researchers used electrical energy to spur the chemical reaction.
In laboratory experiments, a mild electrical current instantaneously changed the structure of marine sand, transforming it into a rock-like, immoveable solid. The researchers are hopeful this strategy could offer a lasting, inexpensive and sustainable solution for strengthening global coastlines.
The study was published Aug. 22 in the journal Communications Earth and the Environment , a journal published by Nature Portfolio.
“Over 40 percent of the world’s population lives in coastal areas,” said Northwestern Engineering’s Alessandro Rotta Loria , who led the study. “Because of climate change and sea-level rise, erosion is an enormous threat to these communities. Through the disintegration of infrastructure and loss of land, erosion causes billions of dollars in damage per year worldwide. Current approaches to mitigate erosion involve building protection structures or injecting external binders into the subsurface.
“My aim was to develop an approach capable of changing the status quo in coastal protection — one that didn’t require the construction of protection structures and could cement marine substrates without using actual cement. By applying a mild electric stimulation to marine soils, we systematically and mechanistically proved that it is possible to cement them by turning naturally dissolved minerals in seawater into solid mineral binders — a natural cement.”
Rotta Loria is the Louis Berger Assistant Professor of Civil and Environmental Engineering at the McCormick School of Engineering. Andony Landivar Macias, a former PhD candidate in Rotta Loria’s laboratory , is the paper’s first author. Steven Jacobsen , a mineralogist and professor of Earth and planetary sciences in Northwestern’s Weinberg College of Arts and Sciences , also co-authored the study.
Sea walls, too, erode
From intensifying rainstorms to rising sea levels, climate change has created conditions that are gradually eroding coastlines. According to a 2020 study by the European commission’s Joint Research Centre, nearly 26 percent of the Earth’s beaches will be washed away by the end of this century.
To mitigate this issue, communities have implemented two main approaches: building protection structures and barriers, such as sea walls, or injecting cement into the ground to strengthen marine substrates, widely consisting of sand. But multiple problems accompany these strategies. Not only are these conventional methods extremely expensive, they also do not last.
By applying a mild electric stimulation to marine soils, we systematically and mechanistically proved that it is possible to cement them by turning naturally dissolved minerals in seawater into solid mineral binders — a natural cement.
Alessandro Rotta Loria Louis Berger Assistant Professor of Civil and Environmental Engineering
“Sea walls, too, suffer from erosion,” Rotta Loria said. “So, over time, the sand beneath these walls erodes, and the walls can eventually collapse. Oftentimes, protection structures are made of big stones, which cost millions of dollars per mile. However, the sand beneath them can essentially liquify because of a number of environmental stressors, and these big rocks are swallowed by the ground beneath them.
“Injecting cement and other binders into the ground has a number of irreversible environmental drawbacks. It also typically requires high pressures and significant interconnected amounts of energy.”
Turning ions into glue
To bypass these issues, Rotta Loria and his team developed a simpler technique, inspired by coral and mollusks. Seawater naturally contains a myriad of ions and dissolved minerals. When a mild electrical current (2 to 3 volts) is applied to the water, it triggers chemical reactions. This converts some of these constituents into solid calcium carbonate — the same mineral mollusks use to build their shells. Likewise, with a slightly higher voltage (4 volts), these constituents can be predominantly converted into magnesium hydroxide and hydromagnesite, a ubiquitous mineral found in various stones.
When these minerals coalesce in the presence of sand, they act like a glue, binding the sand particles together. In the laboratory, the process also worked with all types of sands — from common silica and calcareous sands to iron sands, which are often found near volcanoes.
“After being treated, the sand looks like a rock,” Rotta Loria said. “It is still and solid, instead of granular and incohesive. The minerals themselves are much stronger than concrete, so the resulting sand could become as strong and solid as a sea wall.”
While the minerals form instantaneously after the current is applied, longer electric stimulations garner more substantial results. “We have noticed remarkable outcomes from just a few days of stimulations,” Rotta Loria said. “Then, the treated sand should stay in place, without needing further interventions.”
Read more about Alessandro Rotta Loria
Simple New Process Stores CO2 in Concrete without Compromising Strength
Carbonated concrete developed by Rotta Loria offers the potential to offset emissions from cement manufacturing.
The Ground Is Deforming, and Buildings Aren’t Ready
A study from Rotta Loria, for the first time, linked underground climate change to the shifting ground beneath urban areas.
Opportunities Underground
Rotta Loria’s research explores opportunities and innovations in the subsurface.
Ecofriendly and reversible
Rotta Loria predicts the treated sand should keep its durability, protecting coastlines and property for decades.
Rotta Loria also says there is no need to worry negative effects on sea life. The voltages used in the process are too mild to feel. Other researchers have used similar processes to strengthen undersea structures or even restore coral reefs. In those scenarios, no sea critters were harmed.
And, if communities decide they no longer want the solidified sand, Rotta Loria has a solution for that, too, as the process is completely reversible. When the battery’s anode and cathode electrodes are switched, the electricity dissolves the minerals — effectively undoing the process.
“The minerals form because we are locally raising the pH of the seawater around cathodic interfaces,” Rotta Loria said. “If you switch the anode with the cathode, then localized reductions in pH are involved, which dissolve the previously precipitated minerals.”
Competitive cost, countless applications
The process offers an inexpensive alternative to conventional methods. After crunching the numbers, Rotta Loria’s team estimates that his process costs just $3 to $6 per cubic meter of electrically cemented ground. More established, comparable methods, which use binders to adhere and strengthen sand, cost up to $70 for the same unit volume.
The applications of this approach are countless. Alessandro Rotta Loria
Research in Rotta Loria’s lab shows this approach also can heal cracked structures made of reinforced concrete. Much of the existing shoreside infrastructure is made of reinforced concrete, which disintegrates due to complex effects caused by sea-level rise, erosion and extreme weather. And if these structures crack, the new approach bypasses the need to fully rebuild the infrastructure. Instead, one pulse of electricity can heal potentially destructive cracks.
“The applications of this approach are countless,” Rotta Loria said. “We can use it to strengthen the seabed beneath sea walls or stabilize sand dunes and retain unstable soil slopes. We could also use it to strengthen protection structures, marine foundations and so many other things. There are many ways to apply this to protect coastal areas.”
Next, Rotta Loria’s team plans to test the technique outside of the laboratory and on the beach.
The study was supported by Northwestern’s Center for Engineering Sustainability and Resilience .
Get our news in your inbox.
Sign up for our newsletter.
Check out our magazine.
Find more in depth stories and get to know Northwestern Engineering.
IMAGES
VIDEO
COMMENTS
How To Set Up A Beach Erosion Model. STEP 1: Add about 5 cups of sand to one side of your pan. You will want to build it up on a slope so that when water is added some of the sand is higher. STEP 2: Place some rocks or shells in the sand for a beach theme! STEP 3: Fill a small bottle with water, add a drop of blue food coloring, shake and pour ...
Using sugar cubes, students experiment to see how weathering and sand are related. 4. Building Beaches. In the Building Beaches activity, students model a beach using a small tray filled with sand and water and investigate erosion and how beach formations like headlands appear.
Coastal erosion is the loss or displacement of land from processes such as waves and other actions that remove sand, sediment, and bedrock. This ocean experiment allows you to visualize coastal erosion. This experiment uses water to create erosion but wind is also another contributor to coastal erosion. Storms, currents, and even ice are other ...
Let's find out in this erosion experiment. In some parts of the world, houses are still made from mud and sand, which makes them easier to knock down than homes made of brick and stone. To understand how these materials can (and should) be used to make homes, you'll need to explore how they stand up to the elements. Download Project.
Continual erosion of the shoreline by waves also changes the beach over time. One change that erosion can cause is the appearance of a headland. This is land that juts out from the coastline and ...
Beach erosion often occurs when storms bring great waves ashore to loosen the sandy soil of the coastline and wash it into the sea. Florida's beaches were stable for thousands of years. While storms may have taken some coastal materials seaward, the waves also brought sand ashore. Nature's balance of forces created and sustained our famous ...
Erosion occurs when water washes away dirt, rock, or sand. Erosion happens everyday during man-made events, rainfall, and in the ocean. You have probably seen how people try to defend against erosion by placing large rocks near the edge of a river, lake, or the ocean to try to stop water from washing away the land.
An artistic impression of how electricity could be used to strengthen coastlines. Credit: Northwestern University. A new method developed by Northwestern University uses electrical currents to solidify marine sand, creating durable, rock-like structures that could replace costly traditional coastal defenses like sea walls.. Researchers from Northwestern University have demonstrated that a zap ...
The erosion of rock formations in the water, coral reefs, and headlands create rock particles that the waves move onshore, offshore, and along the shore, creating the beach. Continual erosion of the shoreline by waves also changes the beach over time. When larger and stronger waves hit the shoreline, such as in a storm, more shoreline is eroded.
Erosion in a Bottle Activity Source: Source: Soil Science Society of America. Adapted with permission. Soil erosion is the process of moving soil by water or wind — this happens naturally or through human interference. Preventing soil erosion is important because nutrients are lost, and sediment that accumulates in waterways impacts life there.
What is the difference between weathering and erosion? And what is deposition? Discover the definitions and impacts of all three of these important processes...
Forces of nature like wind, rain, and ice can all cause soil erosion. In this experiment, you'll see how these natural patterns weather your soil landscape in different ways. Materials. Steps. In each container, pour soil up against one side, creating a sloped pile. In the first container, simulate water erosion by spraying water over the soil.
In their laboratory experiments, the team applied mild electrical currents (ranging from 2 to 4 volts) to seawater-saturated sand. The results, published in Communications Earth & Environment, were nothing short of miraculous.The electricity triggered chemical reactions that converted naturally occurring ions and minerals in the seawater into solid calcium carbonate—the same stuff seashells ...
Procedure: 1. Fill the top two-thirds of a paint pan liner with sand. The sand we used is for hermit crabs from the pet department at Walmart, however regular sand is ideal. 2. Press the sand to form a "shoreline". The sand should be about 2 or 3 inches deep. 3. Gently pour water into the pan filling the bottom half.
This Demonstration shows weathering, erosion, and deposition in action as you pour water over the sand in a simple yet effective classroom demonstration. Wat...
The erosion of marine sediments is a pressing issue for coastal areas worldwide. ... R. G. Beach Nourishment: ... Skafel, M. G. & Bishop, C. T. Flume experiments on the erosion of till shores by ...
Stream table experiment TIPS. These stream tables are obviously much shorter than the typical ones you can order from science suppliers. That limits the amount of erosion that can take place. Here are some tips to help them work the best: Start with wet sand, and pack it down tight. If the sand is dry, it will all run down without creating a ...
The organizations' second field experiment - in which biodegradable bladders filled with sand were submerged to accumulate sand - kicked off in the Maldives in October 2019.
1. Place tightly packed clumps of sand, dirt, and grass/dirt into the foil pan. 2. Add water to a measuring cup. Make your predictions. Which pile - sand, dirt, or grass/dirt - will stand up to erosion the best? 3. Gently and slowly pour water over each of the piles. 4.
In laboratory experiments, a mild electrical current instantaneously changed the structure of marine sand, transforming it into a rock-like, immoveable solid. The researchers are hopeful this strategy could offer a lasting, inexpensive and sustainable solution for strengthening global coastlines.
Erosion K-12 experiments & background information for lesson plans, class activities & science fair projects for elementary, middle and high school students. ... Another type of erosion is called Decomposition and is when waves wash away sand and other material from cliffs or beaches. Soil erosion can be prevented several ways.
Sand Erosion. In this hands-on experiment, students will utilize sandpaper to imitate wind erosion on a soft rock like limestone, calcite, or a similar stone. They can compare the original to the new "sanded-down" version to complete the scientific analysis. Learn More: Berkley. 3. Weathering, Erosion, or Deposition Sorting Activity
A tray of sand, water, and some pebbles is all you need to build a working model of coastal erosion. With this experiment, students can see exactly how the smallest movements of water cause significant erosion. Learn More: Little Bins for Little Hands. 13. Try a Chemical Weathering Experiment. This experiment has students discovering how ...
Erosion is one of the slowest, yet most powerful forces in nature. The immensity of the Grand Canyon is an extreme example of the effects erosion can have on its surroundings. ... Another experiment is to place sand in a slope at one side of a pan, and then add enough water to cover about half of it. Using a ruler, create waves that move ...
Soil Experiment #1. Gather up a handful of four different types of soil. This can be dirt, sand, clay, silt, chalk, etc. Place the soil types in different sections of your tray. Soil can be dry or add a little bit of water to build a landform. Draw a picture of your soils and describe the texture, color, and shape of each.
In a recently published study, researchers from the University of California, Santa Cruz, and USGS investigate how coral reefs affect shoreline erosion in the Hawaiian Islands. By analyzing decades of coastal data, the study finds that coral reefs play a crucial and complex role in coastal dynamics, offering significant protection during storm events while reducing long-term erosion in most cases.
Erosion Experiment. This experiment looks at 3 types of erosion: water, wind, and glacial. Erosion occurs when the Earth's surface is worn away. The Grand Canyon was formed by water erosion. Can wind and glacial erosion make a big impact, too? Materials Needed: 3 containers of soil (shoe box size works well) Spray bottle of water; Straw; Ice
With this science experiment, discover the effects of water runoff from different soil conditions. Soil erosion can be caused by waves, wind, water, & more. With this science experiment, discover the effects of water runoff from different soil conditions. Call Now to Join Your Solar Farm: 1-866-541-4177.
Direct impact erosion experiments forUNS N06625 cladding and UNS S32750 were conducted using sand particles carriedby air at ambient temperature and pressure. The sand particle sizes ranged from27 µm to 619 µm, sand particle velocities from about 25 m/s to 160 m/s, andimpact angles from 15° to 90°.
Acidic contaminants are one of the major pollutants that penetrate the soil through various pathways, such as breaking pipes, factory effluents, and chemical leakages. In this study, systematic static and cyclic simple shear tests were performed to explore the behavior of marine sands subjected to acidic conditions, considering the influences of the different cyclic stress ratios and acid ...