Untamed Science

  • Plant Biology
  • Human Biology
  • Biology Plant Biology Phototropism Experiment

Experiment: How do plants “see” light?

Can plants really “see” the light.

Scientists call a plant’s ability to bend toward light phototropism . Even as far back as ancient Greece it’s been a big puzzle about how plants are able to do it. People experimented with how plants accomplish this amazing feat, but no one really figured out how it worked—until Charles Darwin came along, that is.

Although Darwin is most well-known for his studies on evolution, he was also a prolific scientist in general. The questions about phototropism piqued his curiosity, and he thought of an ingenious experiment to test how plants are able to see light. In this experiment, we’ll recreate what he did, and at the end we’ll dive further into the science.

Darwin's Phototropism experiment cartoon

3 small cups full of soil Tape, a marker, and 3 sticky notes Medium-sized box (such as a shoebox or a storage cube) 12 corn seeds Aluminum foil Small cookie sheet that fits inside the box (or another sheet of aluminum foil) 1 Straw Water

  • Plant four corn seeds in each of the soil cups. Make sure they’re evenly spaced, and plant them just a half inch under the dirt.
  • Water the cups, and dump out any excess water (be careful not to tip the soil and seeds out). Place the cups on the cookie sheet or aluminum foil. This will prevent moisture and dirt from soaking through the box.
  • Place the cups/cookie sheet setup inside of the box. Make sure it’s open on one side so that light is coming in from an angle. Place in a windowsill, with the open side facing the sun. (You might need to stack some books underneath it to support it, if your windowsill isn’t very wide.)
  • Shoot cap: Cut a small 2″ x 3″ square of aluminum foil. Wrap it around the tip of a straw to create a small, closed-ended metal cap, and slide it off. This will be placed over the tip of the growing shoot to cover any light coming in to the tip.
  • Base sleeve: Cut a small 1/2″ x 3″ square of aluminum foil. Wrap it around the middle of a straw so it creates a small open-ended 1/2″ tall tube, and slide it off. This will be placed around the growing shoot so that it can grow through it.
  • Check the cups each day. Once they send up a shoot about half an inch high, place either a shoot cap (on Tip seedlings) or a base sleeve (on Base seedlings) around them, depending on which cup they’re in. The control cup will get neither of the light exclusion devices. The seedlings might grow at different rates, so be sure to check each day to put the caps/sleeves on as needed. They grow fast once they germinate!
  • Continue to water the seedlings as needed.
  • Check the seedlings after a week. What has happened? Compare the seedlings with the caps and the sleeves to the control seedlings. Are any of them growing in certain directions?

How did the seedlings “see” the light?

If the experiment worked correctly, you should have noticed that the seedlings that were covered with caps at the tip grew straight up, while the control seedlings and the seedlings with the bases covered bent towards the light. This is phototropism in action.

Darwin correctly concluded that plants are able to “see” light using the tips of the plant shoots, rather than through the stalks. It wasn’t until a bit later that scientists figured out exactly why that was, though.

It turns out that plants are able to grow by using hormones such as auxins and gibberellins . Auxin in particular tells individual cells to reach out and grow longer, like Stretch Armstrong. It’s one of the ways that plants grow taller. Normally, plants growing with an unshaded light source will grow straight up towards the sun because auxin is evenly distributed all around the shoot.

But when the light is heavily shaded and comes in from an angle, something interesting happens. Auxin starts to concentrate on the shaded side of the plant instead, and as a result, the cells on the sunny side stay the same size but the cells on the shaded side grow longer. This causes the plant to tip and grow towards the light.

phototropism - auxin

Auxin is primarily produced in the tips of the plants. This is why the plant grew straight up when you covered the tip with a cap—it couldn’t “see” the light anymore! The tips of the control seedlings and the seedlings with the bases covered could still sense the light, so they grew towards the sunlight.

Thanks to Charles Darwin and modern science, the mystery of how plants grow towards light was finally solved.

Learn more about phototropism:

To understand plant tropisms, you first have to understand plant hormones. We created an excellent page about Plant Growth Hormones  here and here .

Related Topics

Choose one of the following categories to see related pages:.

  • Experiments ,

Share this Page

' src=

Lindsay graduated with a master’s degree in wildlife biology and conservation from the University of Alaska Fairbanks. She also spent her time in Alaska racing sled dogs, and studying caribou and how well they are able to digest nutrients from their foods. Now, she enjoys sampling fine craft beers in Fort Collins, Colorado, knitting, and helping to inspire people to learn more about wildlife, nature, and science in general.

Plant Biology-boo

  • Monocots vs Dicots Explained
  • Phototropism Experiment
  • Plant Diversity
  • Plant Growth Hormones
  • Plant Memory
  • Transport in Plants
  • Water Transport Experiment
  • Why leaves change color

Science Newsletter:

Full list of our videos.

experiment for phototropism

Teaching Biology?

experiment for phototropism

How to Make Science Films

experiment for phototropism

Read our Wildlife Guide

experiment for phototropism

Choose an Account to Log In

Roly

Notifications

Science project, plant phototropism experiment.

experiment for phototropism

As plants grow, they move up toward the light. But what is a plant’s favorite color? Do plants move toward some colors more than others?

Do plants bend toward certain colors of light?

  • 2 1-foot tall cardboard boxes with lids
  • Piece of cardboard
  • 2 small lamps
  • 2 full spectrum light bulbs
  • Box cutter knife
  • Masking tape
  • 1 3” x 3” piece of clear, red, green, and blue cellophane
  • Spray bottle
  • 8 bean seeds
  • 8 small pots
  • First, get your plants growing. Plant two of your bean seeds in two different pots, water them, and wait for them to poke out of the ground.
  • While you’re waiting, get your boxes ready.  Cut a hole 2” in diameter about 3 inches from the bottom of each box. Place the clear cellophane over the hole. This will let all of the light into the box. Over the hole in the other box, place the red cellophane. This will only let red light into the box.
  • Put one plant in the first box and one in the second. Use a ruler to position each bean plant two inches away from the cellophane window.  Take a photo of the plants, looking downward from the top of the box.
  • Put the boxes on different sides of the same room.
  • Now it’s time to light things up! Put the lamps next to the boxes on the side with the cellophane window. Take out your ruler again and measure to make sure that the lamps are the same distance from the hole.
  • Put the lids on each box.
  • Every morning, turn on each lamp. Every night, turn off the lamps before you go to bed. Leave the plants to grow for a week.
  • After a week has passed, remove the lid and take a photo looking downward. Then remove the plants and take a photo from the front. Do the plants look different? Is one taller than the other? Is one twisted in a different direction?
  • Do the same experiment with new bean plants, but change the color of cellophane to blue. Finally, repeat the experiment with green cellophane.
  • Compare the photos of each bean plant after it had been growing for a week. Did the plants turn more toward a certain color? Was there a color they didn’t like?

The control plants will do better than the plants that are only exposed to one wavelength of light. The plants will grow better in red and blue light than in green light. The plants will grow toward red and blue light but will not move toward the green light.

Plants love the light, right? Yes and no. Plants do love the light, but they like some wavelengths of light more than others.

When you look at a rainbow, you can see that the visible spectrum of light actually has different colors or wavelengths inside it.  The visible spectrum is the light that we can see. Different objects reflect different types of light. A blue bowl reflects blue light. A green plant reflects green light.

Inside a plant are chloroplasts . Inside the chloroplasts are tiny molecules called photopigments . Photopigments help the plant absorb light. A plant has different types of photopigments so it can absorb different colors of light.

When natural light shines on a plant, that plant takes in the light from the different wavelengths and uses it to make food.  This natural light is called white light, and it contains all of the types of light. If there’s only one color of light shining on a plant, then only some of the photopigments work, and the plant doesn’t grow as well. This is why your plant under the full light spectrum grew better than the plants with the cellophane filters.

Plants also move toward the light. Seeds push little leaves up from the ground into the light. A house plant in a dark room will grow toward the light. This movement in response to light is called phototropism . When a plant moves toward the light, it’s called positive tropism . When a plant moves away from light, it’s called negative tropism .

How do plants move? They do so with the help of chemicals called auxins . Think of auxins as an elastic band for cells. They help cells get longer and move. Sunlight reduces auxin, so the areas of the plant that are exposed to sunlight will have less auxin. The areas on the dark side of the plant will have more auxin. That means that they will have long, stretchy cells. This allows the plant to move toward the light.

The plants in your experiment likely showed positive tropism, except when it came to the green light. Why did the plants not move toward the green light? Plants are green, which means that they reflect green light. It bounces off the leaves. This means that they can’t use green light very well, and the green light bounces off the plant instead of encouraging movement toward the light.

Digging Deeper

What would happen if you left plants for a long time in light that was only red or blue? Would they survive? 

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.

experiment for phototropism

  • Why Does Water Expand When It Freezes
  • Gold Foil Experiment
  • Faraday Cage
  • Oil Drop Experiment
  • Magnetic Monopole
  • Why Do Fireflies Light Up
  • Types of Blood Cells With Their Structure, and Functions
  • The Main Parts of a Plant With Their Functions
  • Parts of a Flower With Their Structure and Functions
  • Parts of a Leaf With Their Structure and Functions
  • Why Does Ice Float on Water
  • Why Does Oil Float on Water
  • How Do Clouds Form
  • What Causes Lightning
  • How are Diamonds Made
  • Types of Meteorites
  • Types of Volcanoes
  • Types of Rocks

Phototropism

What is phototropism .

‘Photo’ means light, and ‘tropism’ means turning. Phototropism is the movement of plants or other organisms, such as fungi, in the direction of light or away from it. 

experiment for phototropism

Types of Phototropic Movement

Based on the orientation of a plant or other organisms   in response to light, phototropism is of two types:

  • Positive phototropism :   Growing in the direction of a light source, as found in the plant shoot.
  • Negative phototropism (Aphotropism) : Growing away from light, as observed in plant roots .

What Causes Phototropism 

The presence of light is the most important factor in phototropism. Light energy stimulates the rate of photosynthesis in plants causing food production and generation of energy in the form of high-energy phosphate molecule , ATP . Phototropism is thus a survival strategy adopted by plants to maximize their light-harvesting capacity.

Early Experiments to Explain Phototropism

The Experiment of Charles Darwin

In 1880, Darwin and his son Frances in their experiment described the bending of grass seedlings in the presence of light. They used newly grown plants with their shoot tips covered by a sheath called the coleoptile. Through this experiment, they found that coleoptile-covered plants did not respond to the light stimulus. In separate research involving plants where either the tip or the lower part of the coleoptile is covered, they concluded that the upper part is only capable of sensing the light response.

Peter Boysen-Jensen Experiment 

In 1913, Danish physiologist Peter Boysen-Jensen, in a follow-up study to Darwin, showed that a chemical signal produced at the tip of the shoot in plants is responsible for the bending response of the coleoptiles   in response to light. 

In his experiment, he removed the tip of the coleoptile and covered the cut portion with gelatin and then reinserted the tip. To his observation, he found the coleoptile to bend towards the light. On repeating the experiment using mica in place of gelatin, he discovered that the coleoptile lost its bending response. This result proved that gelatin having pore allowed a specific chemical signal to communicate between the tip and the base of the plant while mica being impermeable prevented any such communication. 

In another experiment, Boysen-Jensen further showed that the chemical signal travels only through the coleoptile-covered plant’s shaded side. This proved that the signal is a growth stimulator.

Kenneth Thimann Experiment

Later in 1933, Kenneth Thimann isolated and identified the chemical signal responsible for plant growth and movement as indole-3-acetic acid or auxin . 

Physical Basis of the Phototropic Response

Plants sense light using specialized regions called photoreceptors. Photoreceptors are made of pigment molecules called chromophores that are responsible for absorbing light energy. On absorbing light, the chromophore initiates a signaling pathway that ultimately causes the plant to bend towards or away from the light source. 

How Does Auxin Promote Phototropism

Auxin is the growth hormone produced at the tip of plants. The mechanism showing the role of auxin in phototropic response is described in the following steps:

  • The specific blue-light receptor proteins called phototropins receive the light of wavelength around 450 nanometers and get stimulated to activate several hormones, including the growth hormone, auxin.
  • The phototropin molecules on the bright side absorb lots of light, while molecules on the shady side absorb much less. This causes more auxin to be transported down the shady side, and less to be transported down the bright side. 
  • Auxin stimulates a decrease in cell-pH, which helps to activate a group of enzymes responsible for cell growth and elongation in the shaded regions, causing the stem to bend towards the light source. 

Examples 

Positive Phototropism

  • Sunflower plants  ( Helianthus annus ): Highly phototrophic. They grow towards the sun, but they can also be seen to follow the sun’s movement from east to west. Sunflower plants need more light for growth and survival, as well as for fruiting and flowering, compared to other flowering plants.
  • Seedlings : Grow vertically upwards when the amount of light is the same on all sides. However, when the light is not evenly distributed, the seedling tends to bend in the direction of light.
  • Fungi  (Genus  Pilobolus ): Feed and survive on dead and decaying matter.  Pilobolus crystallinus  (commonly known as the ‘hat-thrower fungus’) uses phototropism to disperse their spores in the vegetation where herbivores consume them. 

Negative Phototropism

  • The plant roots growing downwards and away from the light source shows negative phototropism
  • Promotes food production in plants by photosynthesis
  • Helps in growth of the plant
  • Helps fungi such as  Pilobolus crystallinus  to complete their lifecycle
  • Phototropism Explained – Thoughtco.com
  • Phototropism & Photoperiodism – Khanacademy.org
  • Phototropism – Plantcell.org
  • Phototropism Experiment – Untamedscience.com

Article was last reviewed on Thursday, February 2, 2023

Related articles

Gene Expression

Leave a Reply Cancel reply

Your email address will not be published. Required fields are marked *

Save my name, email, and website in this browser for the next time I comment.

Popular Articles

experiment for phototropism

Join our Newsletter

Fill your E-mail Address

Related Worksheets

  • Privacy Policy

© 2024 ( Science Facts ). All rights reserved. Reproduction in whole or in part without permission is prohibited.

Investigating Phototropism & Geotropism ( OCR A Level Biology )

Revision note.

Alistair

Biology & Environmental Systems and Societies

Investigating Phototropism & Geotropism

Phototropism in plant shoots.

  • Plant shoots are positively phototropic , meaning they grow towards light
  • This ensures they maximise the amount of light they can absorb for photosynthesis
  • Many of the experiments were conducted using coleoptiles (a sheath that surrounds the young growing shoot of grass plants)
  • Darwin discovered that removing the tip of a coleoptile stopped the phototropic response to a unidirectional light source (light coming from one side) from occurring
  • To ensure this was not simply due to the wounding caused to the plant, he covered the tip of a coleoptile with an opaque cover or 'cap' instead, to block out the light. This also stopped the phototropic response from occurring, showing that the tip of the coleoptile was responsible for detecting light
  • Boysen-Jensen found that if he replaced the cut tip back on top of the coleoptile and inserted a gelatin block as a barrier in between, the phototropic response was restored
  • This showed that the stimulus for growth was a chemical (hormone) , which was able to travel through the gelatin block
  • Bosen-Jensen then inserted a mica barrier (mica is impermeable to chemicals) halfway through the coleoptile just below the tip, first on the lit side and then on the shaded side
  • When the mica barrier was inserted into the lit side , the phototropic response occurred
  • When the mica barrier was inserted into the shaded side , the phototropic response did not occur
  • This confirmed that the stimulus for growth was a chemical (hormone) and showed that it was produced at the tip , before travelling down the coleoptile on the side opposite to the stimulus (i.e. the shaded side)
  • It also showed that the stimulus acted by causing growth on the shaded side (rather than inhibiting growth on the lit side)
  • Paál cut off the tip of a coleoptile and then replaced it off-centre in the dark
  • The side of the coleoptile that the tip was placed on grew more than the other side, causing the coleoptile to curve (similar to a phototropic response)
  • This showed that, in the light, the phototropic response was caused by a hormone diffusing through the plant tissue and stimulating the growth of the tissue
  • Went placed the cut tip of a coleoptile on a gelatin block, allowing the hormones from the tip to diffuse into the block
  • The block was then placed on the coleoptile, off-centre and in the dark
  • As in Paál's experiment, the side of the coleoptile that the block was placed on grew more than the other side, causing the coleoptile to curve
  • The greater the concentration of hormone present in the block, the more the coleoptile curved

Historical phototropism experiments 1, downloadable AS & A Level Biology revision notes

Four historical phototropism experiments were conducted to investigate the process by which phototropism occurs

Controlling growth by elongation

  • Indole-3-acetic acid (IAA), which is an auxin , is a specific growth factor found in plants
  • IAA is synthesised in the growing tips of roots and shoots (i.e. in the meristems , where cells are dividing )
  • IAA coordinates phototropisms in plants by controlling growth by elongation
  • IAA molecules are synthesised in the meristem and pass down the stem to stimulate elongation growth
  • The IAA molecules activate proteins in the cell wall known as expansins , which loosen the bonds between cellulose microfibrils , making cell walls more flexible
  • The cell can then elongate

The phototropic mechanism

  • Phototropism affects shoots and the top of stems
  • The concentration of IAA determines the rate of cell elongation within the region of elongation
  • If the concentration of IAA is not uniform on either side of a root or shoot then uneven growth can occur
  • It is described as positive because growth occurs towards the stimulus
  • Experiments have shown that IAA moves from the illuminated side of a shoot to the shaded side
  • The higher concentration of IAA on the shaded side of the shoot causes a faster rate of cell elongation
  • This causes the shoot to bend towards the light

Phototropism mechanism, downloadable AS & A Level Biology revision notes

Higher concentrations of IAA on the shaded side increases the rate of cell elongation so that the shaded side grows faster than the illuminated side

Geotropism in plant shoots and roots

  • Gravity affects both plant shoots and roots , but in different ways
  • Gravity modifies the distribution of IAA so that it accumulates on the lower side of the shoot
  • As seen in the phototropic response, IAA increases the rate of growth in shoots , causing the shoot to grow upwards
  • In roots, higher concentrations of IAA results in a lower rate of cell elongation
  • The IAA that accumulates at the lower side of the root inhibits cell elongation
  • As a result, the lower side grows at a slower rate than the upper side of the root
  • This causes the root to bend downwards

Practical: investigating the effect of IAA on root growth

  • Experiments can be carried out to investigate the effect of IAA on root growth in seedlings
  • Seedlings (of the same age and plant species)
  • Cutting tile
  • Light source
  • Lightproof container
  • Blocks of agar (all the same volume)
  • Use the scalpel to cut a 1cm section from the root tip of each seedling
  • Mark the root tips at 2mm marks
  • The water helps to keep the plant tissue alive
  • Remove the ends of the root tips using the scalpel
  • Transfer root cuttings with the end removed to an agar block
  • A uniform light source is present
  • Transfer intact root tips to an agar block
  • A light-proof container is placed over the seedlings to prevent light from entering
  • Apply a directional light source to one side of the root tips
  • Leave all the roots in their treatment conditions for 3 hours
  • Use the 2mm marker lines to determine if growth has occurred
  • Note if the growth has been even on both sides

Results and analysis

  • IAA is synthesised in the root tips so removing them means that no IAA is produced
  • There is no inhibition of cell elongation
  • There is an equal concentration of IAA on both sides of the root tip
  • The inhibition of cell elongation is equal on both sides of the root tip
  • The roots do not grow as long as those in group A due to the presence of IAA
  • There is a greater concentration of IAA on the shaded side
  • This results in greater inhibition of cell elongation on the shaded side
  • So the illuminated side grows at a faster rate
  • The roots bend away from the light – negative phototropism

Limitations

  • Certain genotypes may be more prone to bending or have slightly different sensitivities to IAA
  • If the root is mishandled the marks can be altered, which will affect the results
  • Only general comments can be made about whether there has been even growth on both sides of the roots

IAA effect on roots, downloadable AS & A Level Biology revision notes

The different treatments produce different levels of growth in the root tips. The IAA molecules inhibit cell elongation in roots

You may sometimes see IAA simply referred to as auxin. IAA is a particular type of auxin, which is a more general term for a particular group of plant hormones.

You've read 0 of your 10 free revision notes

Get unlimited access.

to absolutely everything:

  • Downloadable PDFs
  • Unlimited Revision Notes
  • Topic Questions
  • Past Papers
  • Model Answers
  • Videos (Maths and Science)

Join the 100,000 + Students that ❤️ Save My Exams

the (exam) results speak for themselves:

Did this page help you?

Author: Alistair

Alistair graduated from Oxford University with a degree in Biological Sciences. He has taught GCSE/IGCSE Biology, as well as Biology and Environmental Systems & Societies for the International Baccalaureate Diploma Programme. While teaching in Oxford, Alistair completed his MA Education as Head of Department for Environmental Systems & Societies. Alistair has continued to pursue his interests in ecology and environmental science, recently gaining an MSc in Wildlife Biology & Conservation with Edinburgh Napier University.

  • Search Menu
  • Sign in through your institution
  • Advance articles
  • Darwin Reviews
  • Special Issues
  • Expert View
  • Flowering Newsletter Reviews
  • Technical Innovations
  • Editor's Choice
  • Virtual Issues
  • Community Resources
  • Reasons to submit
  • Author Guidelines
  • Peer Reviewers
  • Submission Site
  • Open Access
  • About Journal of Experimental Botany
  • About the Society for Experimental Biology
  • Editorial Board
  • Advertising and Corporate Services
  • Journals Career Network
  • Permissions
  • Self-Archiving Policy
  • Dispatch Dates
  • Journal metrics
  • Journals on Oxford Academic
  • Books on Oxford Academic

Issue Cover

Article Contents

  • ‘The power of movements in plants’: Darwin's lasting legacy to the field of phototropism research

Darwin's ‘influence’: auxin and its role in phototropism

Darwin's vision: phototropin blue light receptors, getting from darwin's vision to his ‘influence’: early phototropin signalling components, power of movement meets origin: phototropism in the field, from darwin to the future: final thoughts.

  • < Previous

Understanding phototropism: from Darwin to today

  • Article contents
  • Figures & tables
  • Supplementary Data

Jennifer J. Holland, Diana Roberts, Emmanuel Liscum, Understanding phototropism: from Darwin to today, Journal of Experimental Botany , Volume 60, Issue 7, May 2009, Pages 1969–1978, https://doi.org/10.1093/jxb/erp113

  • Permissions Icon Permissions

Few individuals have had the lasting impact on such a breadth of science as Charles Darwin. While his writings about time aboard the HMS Beagle , his study of the Galapagos islands (geology, fauna, and flora), and his theories on evolution are well known, less appreciated are his studies on plant growth responses to a variety of environmental stimuli. In fact, Darwin, together with the help of his botanist son Francis, left us an entire book, ‘ The power of movements in plants ’, describing his many, varied, and insightful observations on this topic. Darwin's findings have provided an impetus for an entire field of study, the study of plant tropic responses, or differential growth (curvature) of plant organs in response to directional stimuli. One tropic response that has received a great deal of attention is the phototropic response, or curvature response to directional light. This review summarizes many of the most significant advancements that have been made in our understanding of this response and place these recent findings in the context of Darwin's initial observations.

‘ The power of movements in plants ’: Darwin's lasting legacy to the field of phototropism research

Plants are sessile by nature, and thus to maximize energy production they must rely on their capacity to move directionally, or exhibit tropic responses, in response to directional environmental cues. The way plants respond to stimuli has fascinated humans since the time of Ancient Greece ( Whippo and Hangarter, 2006 ). Although Charles (and son Francis) Darwin's ‘ The power of movements in plants ’ dealt in large part with Darwin's proposal that circumnutation could provide a unifying model to explain directional growth responses in plants ( Darwin, 1880 ), an hypothesis that has been shown to be generally incorrect, this seminal book has provided the foundation for an entire field of study focused on the tropic responses of plants.

‘ The power of movements in plants ’ proposed several key elements that shape the current research on tropic responses. Darwin, although not the first to do so (for an excellent historical literature review on tropic response research, see Whippo and Hangarter, 2006 ), proposed that plants could grow differentially (thus directionally) in response to external stimuli such as light or gravity. Second, he demonstrated that the part of the plant that perceives the stimulus is separate and distinct from the part that responds to that stimulus. In the case of phototropism, directional light is perceived in the apical portion of a young seedling and ‘transduced’ to more basally localized portions of the shoot as a differential signal that informs the plant which side is the closest to and which is the furthest from the light source such that a bending response occurs ( Fig. 1 ). Finally, Darwin proposed that an ‘influence’ (though he was unable to identify it) moves from the site of stimulus perception to the area of response where bending occurs ( Fig. 1 ).

Stylized representation of phototropism. The diagrams depicted here are meant to represent the coleoptile of a dark-grown grass seedling, such as oat, which is a classic model for phototropism. On the left is a seedling soon after exposure to unidirectional blue light (half sun to the right of the seedling). The ‘eye’ is used to represent Darwin's proposal that light sensing occurred within a specific region of the seedling (in this case the tip of the coleoptile), not that Darwin was ascribing any anthropomorphic properties to the plant per se . The downward pointing black arrows represent the downward flow of the ‘influence’ (we now know as auxin) that Darwin proposed moved from the site of light perception to the site of differential growth that results in curvature. The horizontal black arrows reflect the lateral movement of auxin occurring from the lit to the shaded side of the seedling that has been demonstrated to occur in a variety of plants. On the right is a seedling that has developed a curvature response after a refractory period ( t x ) during which the differentially accumulated auxin (green shading) promotes localized growth (plus signs).

Just after the turn of the century Boysen-Jensen (1911) was able to gain further insight into Darwin's ‘influence’ in an experiment that used pieces of mica to disrupt the proposed influence's flow, and the results of those experiments confirmed that the ‘influence’ does indeed participate in a plant's response to directional stimuli, such as light. In particular, Boysen-Jensen's experiments suggested that Darwin's ‘influence’ flows from the tip of the plant toward the base in the unlit side of the plant, and that this directional and differential movement of the ‘influence’ is critical for the plant's bending response.

Although textbooks generally credit the Dutch plant physiologist Fritz Went with the identification on Darwin's ‘influence’ as the now well-understood plant hormone auxin, the actual history of auxin's chemical identification is a bit more complicated. First, it is important to give shared credit for the physiological identification of auxin to the Russian plant physiologist Nicolai Cholodny, who, while Went was working with grass coleoptiles ( Went, 1926 ), was generating similar results with grass roots ( Cholodny, 1927 ). It is also critical to note that, in actuality, it is Kogl and colleagues ( Kogl and Haagen-Smits, 1931 ) at Utrecht University (where Went did his graduate work) that deserve credit for the first chemical identification of an ‘auxin’ from human urine. Cholodny (1928) and Went (1928) each independently proposed a similar mechanism by which auxin could mediate tropic responsiveness, which was later simply renamed the Cholodny–Went theory ( Went and Thimann, 1937 ). In brief, the Choldony–Went theory combines Darwin's hypotheses with those of the auxin pioneers to propose that an asymmetric accumulation of auxin occurs in response to a tropic stimulus, and that this asymmetric gradient of auxin stimulates the differential growth response that results in tropic curvature. While other models have been proposed, the Cholodny–Went theory is still the prominent one used to explain a plant's response to tropic stimuli.

Following the initial proposal of the Cholodny–Went theory, a number of hypotheses have evolved regarding how the unequal distribution of auxin occurs, particularly in response to phototropic (directional light) stimulation. For example, Went and Thimann (1937) hypothesized that the unequal auxin accumulation occurs due to either light inactivation of auxin on the stimulated side, light-induced inhibition of the production of auxin, or light-induced transport of the auxin from the lit side to the shaded side. A study by Briggs et al. (1957) showed that introducing a physical barrier between the lit and the shaded side of corn coleoptiles disrupts the formation of an auxin gradient, providing evidence against the arguments that light induces the destruction, or the inactivation, of auxin. Subsequently, Briggs (1963) published additional data that provided support for a hypothesis that the unequal distribution of auxin was due to a lateral movement or transport of auxin. Specifically, these data showed that, in maize coleoptiles, an increase in the amount of curvature in response to light was correlated to an increase in the amount of auxin present on the ‘shaded’ side of the coleoptile ( Fig. 1 ).

Pickard and Thimann (1964) applied radio-labelled auxin, indole-3-acetic acid (IAA) in particular, to maize coleoptiles to trace the path of auxin during phototropism. It was found that IAA moves laterally across the coleoptile from the lit to the shaded side under both the pulse (first positive) and extended (second positive) irradiation conditions ( Fig. 1 ). Based on similar radio-tracer labelling studies, Shen-Miller and Gordon (1966) proposed that light promotes a lateral accumulation of auxin by inhibiting polar auxin transport. Gardner et al. (1974) obtained additional support for the notion that light stimulates the lateral movement of auxin, although their data did not support a role of light-mediated inhibition of polar auxin transport.

In recent years, genetic studies in the model plant Arabidopsis thaliana have identified proteins that appear to function as auxin transport facilitators ( Leyser, 2006 ). At least five auxin transport proteins have been associated with stem/shoot phototropism: AUX1 (AUXIN-RESISTANT 1; Stone et al. , 2008 ), PIN1 (PIN-FORMED 1; Blakeslee et al. , 2004 ), PIN3 ( Friml et al. , 2002 ), MDR1 (MULTIDRUG-RESISTANT 1), and PGP1 (P-GLYCOPROTEIN 1; Noh et al. , 2003 ). Studies in Arabidopsis have also led to important findings about how a gradient of auxin established by such transport facilitators leads to differential growth. For example, semi-dominant loss-of-function mutations in the NPH4 (NON-PHOTOTROPIC HYPOCOTYL 4) / ARF7 (AUXIN RESPONSE FACTOR 7) locus, and dominant gain-of-function mutations in MSG2 (MASSUGU 2)/IAA19 and AXR5 (AUXIN-RESISTANT 5) / IAA1 result in severely impaired phototropic and gravitropic responses ( Liscum and Briggs, 1996 ; Watahiki and Yamamoto, 1997 ; Stowe-Evans et al. , 1998 ; Harper et al. , 2000 ; Park et al. , 2002 ; Tatematsu et al. , 2004 ; Yang et al. , 2004 ). NPH4/ARF7 is a transcriptional activator whose activity is repressed in the presence of the MSG2/IAA19 and AXR5/IAA1 ( Liscum, 2002 ). In the presence of elevated levels of free auxin, MSG2/IAA19 and AXR5/IAA1 are rapidly degraded by a 26S proteasome that requires the SCF TIR1 complex containing the auxin receptor TIR1 to target these proteins for degradation ( Tan et al. , 2007 ). This, in turn, allows homodimerization of the NPH4/ARF7 protein and transcription of ‘auxin responsive genes’ ( Tatematsu et al. , 2004 ; Celaya et al. , 2009 ). A recent transcript profiling study in Brassica oleracea has identified a number of genes that appear to represent targets of NPH4/ARF7 regulation in response to tropic stimulation; these include genes encoding proteins involved in the regulation of free auxin levels, additional transcriptional regulators, and proteins involved in the regulation of cell wall extensibility ( Esmon et al. , 2006 ).

In addition to proposing the existence of a mobile ‘influence’ that was necessary for tropic responses (we now know this ‘influence’ to be auxin; see above), Darwin made observations, again presented in ‘ The power of movements in plants ’, that indicated that tropic curvatures in response to light were not general light responses but specific with respect to light quality. In particular, Darwin was able to demonstrate that the blue region of the electromagnetic spectrum is the most effective portion of the spectrum with respect to the induction of phototropism. These findings have, like those of Darwin's tropic ‘influence’, provided the impetus for a large number of studies over the past 100 or so years. Yet only within the past decade or so have the molecular details of how plants ‘see’ blue light cues ( Fig. 1 ) to induce phototropism, ‘Darwin's vision’ if you will, have become known.

As was the case with the elucidation of the molecular mechanisms underpinning the role of auxin in phototropism, Arabidopsis genetics was also a major factor in the identification of the photoreceptor molecules mediating phototropism in higher plants. The first of these photoreceptors identified at the molecular level is phototropin 1 (phot1) ( Huala et al. , 1997 ), originally designated NPH1 (for its n on- p hototropic h ypocotyl mutant phenotype; Liscum and Briggs, 1995 ). The PHOT2 gene was subsequently identified based on its high degree of sequence homology to PHOT1 ( Jarillo et al. , 2001 ; Sakai et al. , 2001 ). Phototropins regulate not just phototropism, but a number of additional blue light responses, including stomatal opening, chloroplast movements, leaf movements and expansion, and rapid inhibition of stem growth ( Christie, 2007 ).

The functions of the phototropins in these responses are both overlapping and distinct. For example, in the case of phototropism, phot1 (see Briggs et al. , 2001 , for a description of nomenclature) is the dominant receptor, mediating response across a wide range of fluence rates (e.g. 0.01–100 μmol m -2 s -1 ), whereas phot2 appears to operate only at higher fluence rates (>10 μmol m -2 s -1 ) ( Sakai et al. , 2001 ). By contrast, with respect to stomatal regulation, both phot1 and phot2 contribute over the entire range of effective fluence rates ( Kinoshita et al. , 2001 ; Kinoshita and Shimazaki, 2002 ). The interplay between the phototropins is even more complex when one considers blue light-induced chloroplast movements. In high-light conditions, chloroplasts move away from the upper surface of the leaf to avoid photobleaching ( Wada et al. , 2003 ), a response that is mediated solely by phot2 ( Jarillo et al. , 2001 ; Kagawa et al. , 2001 ). However, in low light, both phot1 and phot2 appear to contribute equally to the accumulation of chloroplasts along the upper surface of the leaf to maximize photosynthetic light capture ( Wada et al. , 2003 ).

PHOT1 and PHOT2 , being duplicate genes, encode proteins that are strikingly similar in their overall sequence and structure ( Christie, 2007 ). Structurally, the phototropins consist of two major parts: (i) an amino-terminal photosensory domain, and (ii) a carboxyl-terminal Ser/Thr protein kinase signalling domain ( Fig. 2 ). Both portions of the protein are necessary for phototropic signalling and much has been learned in recent years about how each portion functions and is regulated, as discussed below.

Domain organization of the phototropin blue light receptors. Both phototropins (phot1 and phot2) share the same basic organization with two amino-terminal LOV ( l ight, o xygen, and v oltage) domains and a carboxyl-terminal protein kinase domain. Although not shown here, a single molecule of FMN (flavin mononucleotide) is associated with each LOV domain as a light-harvesting cofactor.

LOVing blue light: photosensory mechanism of phototropins

The photosensory domain of a phot contains two ∼110 amino acid islands with homology to each other that are critical for photoreceptor activity ( Christie, 2007 ): LOV1 (light, oxygen, voltage) and LOV2 ( Fig. 2 ). The LOV domains are members of the larger PAS (Per, Arnt, Sim) domain superfamily ( Huala et al. , 1997 ; Crosson et al. , 2003 ). Each of the LOV domains binds a single molecule of blue light-absorbing flavin mononucleotide (FMN) ( Christie et al. , 1998 ), imparting photoreceptor function to the phototropins ( Christie et al. , 1999 ; Salomon et al. , 2000 ).

As shown in Fig. 3 , photosensitive LOV domains undergo a unique photocycle in response to absorption of blue light ( Celaya and Liscum, 2005 ; Christie, 2007 ; Matsuoka et al. , 2007 ). In darkness, the FMN chromophore is bound non-covalently to the LOV domain as a singlet ground state molecule. This state, which is capable of absorbing blue light, is referred to as LOV D 447 ( Salomon et al. , 2000 ; Crosson and Moffat, 2001 ; Swartz et al. , 2001 ) and absorption of a single photon of blue light results in the generation of an excited singlet FMN, which is rapidly converted into a red-shifted triplet state (LOV L 660 ) ( Swartz et al. , 2001 ; Corchnoy et al. , 2003 ; Kennis et al. , 2003 ; Kottke et al. , 2003 ). The triplet state flavin rapidly decays to form a covalent adduct between the C(4a) atom of the FMN and the cysteine within a highly conserved motif (GXNR C FLQ) in the LOV domain; a state with a near UV-shifted absorption maximum designated LOV S 390 ( Salomon et al. , 2000 ; Crosson and Moffat, 2001 , 2002 ; Swartz et al. , 2001 ; Kasahara et al. , 2002 ; Fedorov et al. , 2003 ; Kennis et al. , 2003 ; Kottke et al. , 2003 ). This FMN-cysteinyl adduct is completely reversible in darkness ( Salomon et al. , 2000 ; Swartz et al. , 2001 ; Kasahara et al. , 2002 ; Kennis et al. , 2003 ) or after absorption of a second near UV photon ( Kennis et al. , 2004 ). Thus, LOV domains cycle between two major states (with a transient intermediate): the dark state (LOV D 447 ) and the lit state (LOV S 390 ), depending upon the light condition.

Proposed photocycle of the phototropin LOV domains. The photocycle begins with the absorption of a blue photon of light by the dark-state (LOV D 447 ) and subsequent conversion to an excited singlet state (asterisk). The excited singlet state is then converted to the red light-absorbing (LOV L 660 ) excited triplet state (T), which is hence converted into the near-UV-absorbing covalent adduct (LOV S 390 ) that represents the active state. Both the singlet and active LOV S 390 states can be converted to the initial dark-state by incubation in darkness. Details of this photocycle are described in the text. Approximate half-times of reactions are given.

It is generally accepted that the LOV S 390 FMN-cysteinyl adduct represents the active signalling state of a phototropin ( Crosson et al. , 2003 ; Celaya and Liscum, 2005 ; Christie, 2007 ; Matsuoka et al. , 2007 ). Consistent with this model, replacement of the critical cysteine with either serine or alanine eliminates the formation of LOV S 390 ( Salomon et al. , 2000 ; Swartz et al. , 2002 ; Kottke et al. , 2003 ), and expression of a PHOT1 transgene containing the cysteine to alanine mutation in both LOV1 and LOV2, or LOV2 alone, fails to complement the aphototropic phenotype of a phot1 null mutant ( Christie et al. , 2002 ). It is interesting to note that a LOV1 cysteine to alanine single mutant transgene does compliment the aphototropic phot1 mutant phenotype, indicating that the two LOV domains are not equal with respect to physiological function ( Christie et al. , 2002 ; Sullivan et al. , 2008 ). Similarly, LOV1 is dispensable, whereas LOV2 is sufficient on its own to mediate function of phot2 in the chloroplast avoidance response ( Kagawa et al. , 2004 ).

The aforementioned findings raise an obvious question: what is the functional role of the LOV1 domain? Several independent studies ( Salomon et al. , 2004 ; Nakasako et al. , 2004 ; Katsura et al. , 2008 ), suggest that LOV1 may function as a dimerization motif; a finding certainly not at odds with the fact that LOV domains are a sub-class within the larger PAS domain superfamily ( Crosson et al. , 2003 ). It is also interesting to note that the quantum efficiency for conversion of LOV D 447 to LOV S 390 is about 10-fold higher in LOV2 than LOV1 in phot1 ( Salomon et al. , 2000 ; Kasahara et al. , 2002 ; Iwata et al. , 2005 ), although once photoconverted the LOV1 domain is longer-lived than LOV2 ( Kasahara et al. , 2002 ; Iwata et al. , 2005 ). These observations suggest that at least in the case of phot1, the predominant receptor mediating phototropism, the LOV2 domain is considerably more ‘photodynamic’ than LOV1, and that selective pressures in nature are stronger on LOV2 versus LOV1. It remains to be determined why, if it is not functioning to regulate phototropin activity, LOV1 remains photosensitive at all.

Sharing the LOV: protein kinase domain activation

As already mentioned the phototropins contain a Ser/Thr protein kinase domain in their carboxyl-terminal regions ( Christie, 2007 ). While no native substrates for the phot protein kinase domain, other than the phototropins themselves, are currently known ( Christie, 2007 ; Matsuoka et al. , 2007 ), mutational studies have demonstrated that the catalytic activity of this domain is apparently necessary for phototropic signal-output ( Christie et al. , 2002 ; Cho et al. , 2007 ). Because of this latter fact much effort has been focused in recent years on understanding how the blue light-dependent formation of LOV2 S 370 leads to activation of the protein kinase domain.

While initial X-ray crystallography studies suggested that only minimal changes occur in the tertiary structure of a LOV domain during photocycling ( Crosson and Moffat, 2001 , 2002 ; Fedorov et al. , 2003 ), solution spectroscopy provided clear evidence that, in fact, the structural rearrangements associated with the formation of LOV S 370 are fairly pronounced, especially if the polypeptides being examined also encompassed regions carboxyl-terminal to LOV2 ( Swartz et al. , 2002 ; Corchnoy et al. , 2003 ; Eitoku et al. , 2005 ; Iwata et al. , 2005 ). Nuclear magnetic resonance (NMR) studies by Harper and colleagues ( Harper et al. , 2003, 2004 ) identified an alpha-helical region (designated the Jα-helix) that resides between LOV2 and the protein kinase domain, which, in darkness, associates with the solvent-exposed surface of the β-sheet portion of the LOV2 core region facing away from the FMN chromophore. Upon blue light-induced FMN-cystenyl adduct formation the Jα-helix becomes disordered and dissociates from LOV2 ( Harper et al. , 2004 ). A number of mutations were identified that could mimic the aforementioned ‘dissociated state’ in the absence of light exposure, and, when introduced into a full-length phot1, these same mutations resulted in light-independent autophosphorylation of the phot1 protein, suggesting that the LOV2-Jα-helix interaction acts to repress the protein kinase activity of phot1 ( Harper et al. , 2004 ). In support of such a LOV2 domain ‘repression model’ ( Fig. 4 ), in vitro studies have shown that an isolated protein kinase domain from phot2 is catalytically active against casein, a common in vitro substrate for protein kinase assays ( Matsuoka and Tokutomi, 2005 ). These results further suggest that phototropins may target proteins other than themselves for phosphorylation in planta , although again no such substrates are currently known.

‘Repression domain’ model for photoactivation of the phototropin protein kinase domain. (A) In darkness the LOV2 domain and cis -associated Jα helix adopt a compressed closed configuration that appears to repress the activity of the carboxyl-terminal protein kinase domain. (B) Upon absorption of a photon of blue light by the non-covalently associated FMN molecule [conversion of LOV2 D 447 to LOV2 L 660 is represented by the transition from oval FMN in (A) to starred FMN in (B)], the active cysteinyl-LOV2 domain adduct is formed to induce progressive structural changes in the LOV2 [represented by the larger blue LOV domains in (B) as compared to smaller grey LOV domains in (A)] that result in unfolding of the Jα helix and de-repression of the protein kinase domain. The active protein kinase domain can then catalyse the autophosphorylation of phototropin (blue balls with a P in each) and currently unknown substrates. The positions of the autophosphorylation sites are generally representative of those determined experimentally but are not meant to depict precisely and exclusively those sites. (This figure is available in colour at JXB online.)

Moving distances with LOV: intracellular localization of phototropins is dynamic and light regulated

Movements of the phototropins are not limited to the angstrom-level intramolecular movements upon formation of LOV S 390 , rather the entire phototropin protein appears to move from one part of the cell to another in response to blue light. For example, while phot1 is normally tightly associated with the plasma membrane in dark-grown seedlings, probably through its carboxyl-terminal protein kinase domain ( Kong et al. , 2006 ), blue light induces the relatively rapid (within minutes) movement of some proportion of phot1 to intracellular locations ( Sakamoto and Briggs, 2002 ; Wan et al. , 2008 ). Similar relocalization properties have also been observed for phot2; although the intracellular compartment to which phot2 moves appears to be the Golgi ( Kong et al. , 2006 ). At present it is unknown exactly how phototropin movement is linked to a particular physiological response, however, a recent study by Han and colleagues ( Han et al. , 2008 ) suggests that this dynamic response may be coupled with receptor adaptation/desensitization or signal attenuation. Specifically the authors found that blue light-induced relocalization of phot1 can be largely, if not completely, prevented by prior exposure to red light ( Han et al. , 2008 ); light conditions that also lead to phytochrome A-mediated enhancement of phot1-dependent phototropism ( Stowe-Evans et al. , 2001 ; Han et al. , 2008 ). Thus it would appear that plasma membrane-localized phot1 is more ‘active’ in terms of phototropic signalling than internalized phot1. Certainly the dynamic nature of phototropin localization represents fertile ground for future studies.

One of the biggest questions currently facing the community of researchers who study phototropism at the molecular level is: how does phototropin activation lead to auxin-regulated differential growth (curvature)? While the details of this process still remain largely unknown, several components of this ‘black box’ have been identified, most notably three phot1-interacting proteins: NPH3 (NON-PHOTOTROPIC HYPOCOTYL 3; Motchoulski and Liscum, 1999 ), RPT2 (ROOT PHOTOTROPISM 2; Inada et al. , 2004 ), and PKS1 (PYTOCHROME KINASE SUBSTRATE 1; Lariquet et al. , 2006 ).

NPH3 and RPT2 are paralogous proteins that represent the founding members of the moderately sized NRL ( N PH3/ R PT2- L ike) protein family (33 members in total) in Arabidopsis ( Celaya and Liscum, 2005 ; Celaya et al. , 2009 ). Members of the NRL family, including NPH3 and RPT2, share five regions of primary sequence conservation ( Fig. 5 ): DIa ( D omain Ia ) and DIb, together comprising an amino-terminal BTB ( B road-Complex/ T ramtrack/ B ric-à-Brac) domain ( Aravind and Koonin, 1999 ; Stogios et al. , 2005 ); DII, resembling no known structural or functional motif; and the remaining two regions, DIII and DIV, together representing the Pfam ‘NPH3 domain (PF03000)’ of unknown function ( Finn et al. , 2008 ). The NRL family also exhibits several regions of conserved predicted secondary structure that have diverged in sequence ( Fig. 3 ); most notably a carboxyl-terminal coiled-coil ( Lupas and Gruber, 2005 ) that is present in approximately half of the family members ( Celaya and Liscum, 2005 ; Pedmale and Liscum, 2007 ). Although the functional roles of each of the aforementioned ‘domains’ are currently not fully understood, the coiled-coil has been shown to represent the phot1-interaction domain of NPH3 ( Motchoulski and Liscum, 1999 ; Pedmale and Liscum, 2007 ), while the BTB domain can mediate interaction between NPH3 and RPT2, at least in yeast ( Inada et al. , 2004 ). Recent studies have shown that the BTB domain of NPH3 can also mediate interaction with CULLIN 3 (CUL3) ( Pedmale and Liscum, 2007 ). These latter results suggest that NPH3 may represent the substrate adapter component of a CUL3-based E3 ubiquitin ligase, a recently recognized role for many BTB-containing proteins ( Krek, 2003 ; Pintard et al. , 2004 ; van den Heuvel, 2004 ; Willems et al. , 2004 ; Stogios et al. , 2005 ; Perez-Torrado et al. , 2006 ).

Domain organization of the NRL ( N PH3/ R PT2- L ike) family of proteins. Members of the NRL family, including the founding members and phototropic signal transduction components NPH3 and RPT2, share five domains of conserved sequence homology designated DIa to DIV. They also contain two regions of conserved predicted secondary structure; an amino-terminal BTB domain (encompassing most of DIa and DIb) and a carboxyl-terminal coiled-coil (C-C).

While no target for ubiquitination by an NPH3-CUL3 complex has yet been reported, one can imagine that such a target might function in the regulation of auxin transport. Findings that phototropic stimulation fails to induce an asymmetric distribution of auxin across the coleoptile in the rice mutant cpt1 ( c oleoptile p hoto t ropism 1 ) ( Haga et al. , 2005 ) is consistent with this hypothesis. CPT1 encodes the rice orthologue of Arabidopsis NPH3 ( Haga et al. , 2005 ), thus placing NPH3/CPT1 downstream of phot1 and upstream of the regulation of auxin redistribution. Recent studies suggest that regulation of auxin transport may represent a common function of NRL family members. For example, mutations in the NPY/ENP/MAB4 ( N AKED P INS IN Y UC MUTANTS/ E NHA N CER OF P INOID/ MA CCHI- B OU 4 ) subfamily of the NRL superfamily appear to influence auxin-mediated organogenesis through alterations in auxin movement ( Cheng et al. , 2007 ; Furutani et al. , 2007 ), probably through genetic interactions between the NPY proteins and the AGC kinases PID (PINOID), PID2, WAG1 (denotes the ‘wagging’ root growth it mediates), and WAG2 ( Cheng et al. , 2008 ). This latter observation is particularly intriguing as phot1 is also an AGC kinase ( Bögre et al. , 2003 ; Galván-Ampudia and Offringa, 2007 ).

The PKS1 protein was originally identified as a negative regulator of phytochrome signalling and to serve as a substrate for phytochrome's protein kinase activity ( Fankhauser et al. , 1999 ), but has since been shown to function as a positive regulator of phototropism as well and physically to interact with both phot1 and NPH3 ( Lariquet et al. , 2006 ). At present, it is not understood how PKS1 (or PKS2 and PKS4; Lariquet et al. , 2006 ) influences phototropism at a molecular level, but it is tempting to speculate that the PKS proteins may bridge the enhancing influences of phytochrome on phot1-dependent phototropism ( Liscum and Briggs, 1996 ; Parks et al. , 1996 ; Janoudi et al. , 1997 ; Stowe-Evans et al. , 2001 ; Liscum, 2002 ), possibly through influences on phot1 localization ( Han et al. , 2008 ). It is also worth noting that NPH3, like PKS1 ( Fankhauser et al. , 1999 ), is a phosphoprotein whose phosphorylation state and functional activity is light-dependent; whereas red light stimulates the phosphorylation of PKS1 in a phytochrome-dependent fashion ( Fankhauser et al. , 1999 ), blue light results in the desphosphorylation of NPH3 that is dependent upon the presence of phot1 ( Pedmale and Liscum, 2007 ). Thus it would appear that the signalling capacity of both NPH3 and PKS1 are regulated by similar post-translational mechanisms linked to the photoreceptors through which the former molecules signal.

Though Darwin (1880) hypothesized that phototropic responses are adaptive to a plant, and this proposal has been reiterated many times over the past 100 plus years in one form or another ( Iino, 1990 ; Liscum and Stowe-Evans, 2000 ; Christie, 2007 ), it has only been within the last few years that this hypothesis has actually been experimentally tested. Galen et al. (2004) have shown that the fitness of field-grown Arabidopsis plants carrying loss-of-function mutations in PHOT1 are significantly lower than that of wild-type plants grown in the same plots. Somewhat surprisingly, in contrast to previous proposals that stem phototropism would represent the adaptive response in nature ( Iino, 1990 ), this study found that root phototropism was the trait coupled to fitness, and only under high light conditions ( Galen et al. , 2004 ). A subsequent study demonstrated that negative root phototropism (bending away from directional blue light) enhances the ability of the plant to access water, which under high light conditions is more abundant deeper in the soil because of increased evaporation near the surface ( Galen et al. , 2007 a ). Three life history traits in particular were shown to be influenced dramatically by the ability of a root to access water in arid conditions: (i) seedling establishment, (ii) accumulation of biomass in established plants, and (iii) fecundity of plants reaching adulthood ( Galen et al. , 2007 a ). These studies provide an exciting potential avenue to develop plants capable of growing in more arid environment that maintain, or even increase, their production value through genetic engineering of phot1 signalling ( Galen et al. , 2007 b ).

Darwin's ‘ The power of movements in plants ’ undoubtedly stimulated an entire field of study on plant responses to the environment. Since publication of this seminal work our understanding of phototropism in higher plants has expanded tremendously. Several significant findings have been made: identification and characterization of the photoreceptors controlling phototropism; identification of auxin as the major growth regulator involved in the development of phototropic responses, and elucidation of its mechanistic basis of action; identification of several signalling components functioning between photoperception and auxin responsiveness; and characterization of an adaptive significance for phototropism under natural growth conditions. However, the goal remains fully to elucidate all of the molecular components, from the reception of light to movement, that contribute to the phototropic response, as well as the ecological variables that have provided the selective pressures for the evolution of this response in nature. In all these regards, there is much work left to do. The next century, like the last, is likely to bring many answers to such questions, leading to an even greater appreciation of just how important ‘ The power of movements in plants ’ really is!

Google Scholar

Google Preview

Month: Total Views:
December 2016 1
January 2017 66
February 2017 67
March 2017 70
April 2017 26
May 2017 21
June 2017 16
July 2017 17
August 2017 19
September 2017 30
October 2017 33
November 2017 56
December 2017 234
January 2018 213
February 2018 327
March 2018 417
April 2018 495
May 2018 343
June 2018 176
July 2018 158
August 2018 281
September 2018 387
October 2018 364
November 2018 449
December 2018 282
January 2019 241
February 2019 312
March 2019 373
April 2019 250
May 2019 279
June 2019 171
July 2019 282
August 2019 316
September 2019 454
October 2019 337
November 2019 334
December 2019 194
January 2020 178
February 2020 286
March 2020 189
April 2020 333
May 2020 199
June 2020 243
July 2020 154
August 2020 166
September 2020 252
October 2020 269
November 2020 418
December 2020 259
January 2021 295
February 2021 298
March 2021 409
April 2021 405
May 2021 414
June 2021 241
July 2021 260
August 2021 224
September 2021 285
October 2021 382
November 2021 435
December 2021 296
January 2022 265
February 2022 266
March 2022 368
April 2022 418
May 2022 379
June 2022 331
July 2022 162
August 2022 161
September 2022 362
October 2022 406
November 2022 417
December 2022 243
January 2023 250
February 2023 258
March 2023 284
April 2023 199
May 2023 216
June 2023 204
July 2023 99
August 2023 166
September 2023 285
October 2023 318
November 2023 322
December 2023 321
January 2024 316
February 2024 290
March 2024 375
April 2024 313
May 2024 360
June 2024 264
July 2024 142
August 2024 93

Email alerts

Citing articles via.

  • Recommend to your Library

Affiliations

  • Online ISSN 1460-2431
  • Print ISSN 0022-0957
  • Copyright © 2024 Society for Experimental Biology
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Institutional account management
  • Rights and permissions
  • Get help with access
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

  • COVID-19 Tracker
  • Biochemistry
  • Anatomy & Physiology
  • Microbiology
  • Neuroscience
  • Animal Kingdom
  • NGSS High School
  • Latest News
  • Editors’ Picks
  • Weekly Digest
  • Quotes about Biology

Biology Dictionary

Phototropism

BD Editors

Reviewed by: BD Editors

Phototropism Definition

Phototropism is the ability of a plant, or other photosynthesizing organism, to grow directionally in response to a light source.

Phototrophic Stimulus Response

Plants and other autotrophs need to manufacture their own food; they usually do this through photosynthesis . Through photosynthesis, organisms convert water, carbon dioxide (CO 2 ) and light into sugars, which are used for energy and growth.

Plants are sessile , meaning they cannot move around to acquire what they need, so in order to maximize the amount of light that they receive through the leaves, they use phototropism.

Positive phototropism is the response of a plant toward a light source, while negative phototropism (also called “aphototropism ”) causes growth in the opposite direction. Plant roots usually use negative phototropism although additionally they use “ gravitropism ”, which is the response to gravitational pull.

How Does Phototropism Work?

The plants first sense the light using photoreceptors . Photoreceptors are special molecules consisting of a protein and a pigment that absorbs light called a chromophore . When light is absorbed by the chromophore, the protein changes shape, initiating a signaling pathway . Plants use signaling pathways to initiate processes such as gene expression, hormone production and growth. The specific photoreceptors which are responsible for detecting light during phototropism are called phototropins .

In 1880, Charles Darwin and his son Frances, discovered that seedlings with the very tip of a sheath called the coleoptile covered did not respond to light, whereas those with the lower part of the coleoptile covered did. From this they theorized that the light-sensing activity took place within the tips of the plant.

Plants contain a hormone called auxin , which coordinates many growth and behavioral processes throughout their life cycles. It is this auxin, which is responsible for the curvature of the stems, allowing plants to grow in a certain direction.

When the phototropins are activated by a light source, auxins are redistributed up the coleoptile and toward the side of the stem where the phototropins are less active – the shaded side.

The auxin activates proton pumps , which lowers the pH of the cells, making them more acidic. This acidification activates enzymes called expansins , which cause the cell walls to become more flexible by breaking the hydrogen bonds . When cell walls are less rigid, the cell walls are able to grow larger and faster than usual.

The larger size of the cells on the shaded size causes an asymmetry of cell size within the stem, and thus, the stem bends toward the light.

Phototropism diagram

Examples of Phototropism

An example of a plant that is highly phototrophic is the sunflower ( Helianthus annus ).

Not only do sunflowers grow toward the sun, they can be visibly seen to track the sun’s movements from East to West throughout the day. At night, the heads move back from West to East in anticipation of the next day’s sunrise.

Scientists have discovered that sunflowers need more light for growth and survival, as well as fruiting, and flowering, than most other flowering plants. This finding is a likely reason for the daily solar tracking activity.

Pilobolus Fungi

Fungi in the genus Pilobolus are saprobic feeders, meaning they feed from non-living or decaying organic material.

In the case of the species Pilobolus crystallinus (commonly known as the “hat-thrower fungus”), they survive by eating the feces of grazing herbivores.

In order to gain access to feces, P.crystallinus uses an explosive propulsion technique, whereby the spores are shot from the sporangiophore in to the air and attach to vegetation. When grazing animals eat the vegetation, the spores pass through the animal’s digestive system and end up in their feces.

P.crystallinus uses phototropism so that their spores will be directed toward a light space where there is likely to be a gap in the grass, so they have a better chance of dispersal. Animals do not usually eat near dung, so they need to disperse their spores away from the dung on which their mycelium grew to increase their chances of being consumed.

Pilobolus crystallinus

Related Biology Terms

  • Photosynthesis – The process by which photosynthesizing organisms convert light energy, carbon dioxide and water into oxygen and organic compounds.
  • Skototropism – The growth of organisms away from a light source, often utilized by root climbers in order to find structures to attach to.
  • Photoperiodism – The mechanism by which organisms respond to seasonal changes in day length.

Cite This Article

Subscribe to our newsletter, privacy policy, terms of service, scholarship, latest posts, white blood cell, t cell immunity, satellite cells, embryonic stem cells, popular topics, digestive system, endocrine system, water cycle, photosynthesis, horticulture, acetic acid, adenosine triphosphate (atp).

Embibe Logo

Phototropism and Geotropism

To study the phenomenon of phototropism and geotropism in plants

experiment for phototropism

The experiment on phototropism and geotropism focuses on understanding how plants respond to external stimuli, specifically light and gravity, and how these responses impact their growth and orientation. Phototropism investigates how plants bend or grow towards a source of light, emphasising their ability to optimise photosynthesis for energy production. Geotropism, on the other hand, explores how plants respond to gravity by either growing downward (positive geotropism) or upward (negative geotropism), which is crucial for proper root development and nutrient absorption.

This experiment sheds light on the remarkable adaptive mechanisms of plants, demonstrating their capacity to sense and react to their environment, ultimately influencing their overall growth and orientation. 

Here are the steps of the experiment on phototropism and geotropism:

  • Take two test tubes and fill them up to about two-thirds of their height with water. Label one as “Test Tube A” and the other as “Test Tube B.”
  • Insert one plant into each test tube, ensuring that the roots are submerged in the water, while the stems and leaves extend out of the test tubes. Secure the openings of the test tubes with cotton balls to hold the plants in place.
  • Seal the mouth of both test tubes tightly using additional cotton and adhesive tape to prevent any water leakage, even when the test tubes are turned upside down.
  • Set up Test Tube A in an upright position using a laboratory stand, and fix Test Tube B upside down in another laboratory stand, ensuring that no water spills out.
  • Place the laboratory stands near a window to expose the plants to direct sunlight.
  • Over the course of the experiment, which includes days 2, 3, and 4, carefully observe both plants. Record your observations, paying attention to the direction in which the stems and primary roots are growing. Note any features that exhibit positive or negative phototropism and positive or negative geotropism.

In summary, this experiment revealed how plants adapt to their surroundings. We observed how they respond to light and gravity. Stems showed positive phototropism by leaning towards light for better photosynthesis, while roots exhibited positive geotropism by growing downward for stability and nutrient uptake. Stems also displayed negative geotropism, growing away from gravity to find optimal light. Overall, this experiment highlighted plant’s remarkable ability to sense and adapt to their environment, shedding light on their intricate growth mechanisms.

FAQs on Phototropism and Geotropism

Q.1 what is phototropism.

Ans. Phototropism is a plant’s growth response to light. It involves plants bending or growing towards a source of light to optimise photosynthesis.

Q.2 What is geotropism?

Ans. Geotropism, also known as gravitropism, is a plant’s growth response to gravity. It involves the orientation of plant roots downward (positive geotropism) and stems upwards (negative geotropism) due to the influence of gravity.

Q.3 What is the role of auxin in phototropism?

Ans. Auxin is a plant hormone that plays a crucial role in phototropism by promoting cell elongation on the shaded side of a plant, causing it to bend towards the light source.

Q.4 What are the practical applications of studying phototropism and geotropism?

Ans. Understanding these plant responses is essential for agriculture, as it helps optimise plant growth and crop yield. It also aids in designing controlled environments for plant cultivation.

Q.5 Can phototropism and geotropism occur simultaneously in the same plant?

Ans. Yes, phototropism and geotropism can occur simultaneously in a plant. For instance, the stem may bend towards the light source (phototropism), while the roots grow downward into the soil (geotropism).

experiment for phototropism

Experiments Related to this Topic

Esterification reaction, how can we prepare soap in the lab, comparing the foaming capacity of soap samples, effect of mass on period of a simple pendulum, effect of amplitude on period of a simple pendulum, cleaning capacity of soap, oxidation reaction of alcohol with alkaline kmno4, study reaction of zinc with sulphuric acid, complete combustion of alcohol, properties of acetic acid.

Embibe

39 Insightful Publications

World Economic Forum

Embibe Is A Global Innovator

accenture

Innovator Of The Year Education Forever

Interpretable And Explainable AI

Interpretable And Explainable AI

Tedx

Revolutionizing Education Forever

Amazon AI Conclave

Best AI Platform For Education

Forbes India

Enabling Teachers Everywhere

ACM

Decoding Performance

World Education Summit

Leading AI Powered Learning Solution Provider

Journal of Educational Data Mining

Auto Generation Of Tests

BW Disrupt

Disrupting Education In India

Springer

Problem Sequencing Using DKT

Fortune India Forty Under Fourty

Help Students Ace India's Toughest Exams

Edtech Digest

Best Education AI Platform

Nasscom Product Connect

Unlocking AI Through Saas

Tech In Asia

Fixing Student’s Behaviour With Data Analytics

Your Story

Leveraging Intelligence To Deliver Results

City AI

Brave New World Of Applied AI

vccircle

You Can Score Higher

INK Talks

Harnessing AI In Education

kstart

Personalized Ed-tech With AI

StartUpGrind

Exciting AI Platform, Personalizing Education

Digital Women Award

Disruptor Award For Maximum Business Impact

The Mumbai Summit 2020 AI

Top 20 AI Influencers In India

USPTO

Proud Owner Of 9 Patents

Close

Trending Searches

growth

Previous Year Question Papers

Sample papers.

Study Mind logo

Personalised lessons and regular feedback to ensure you ace your exams! Book a free consultation today

100+ Video Tutorials, Flashcards and Weekly Seminars

Gain hands-on experience of how physics is used in different fields. Experience life as a uni student and boost your university application with our summer programme!

  • Revision notes >
  • A-Level Biology Revision Notes >
  • CIE A-level Biology Revision Notes

Experiments to Show Phototropism (A-level Biology)

Experiments to show phototropism, investigating plant responses, phototropism experiment.

We can investigate phototropism in plants using the following method. This will allow us to see the response of plants to light.

  • Use 9 plant shoots. Plant all the shoots in individual plant pots, with the same soil type in each pot. Ensure that all the shoots are roughly the same height.
  • Wrap some of the shoots in foil. Now, wrap the tips of 3 shoots in foil. For another 3 shoots, wrap the base of the shoots in foil. Leave the final 3 shoots without foil.
  • Place the shoots under a light source. Place all 9 shoots under a light source for 2 days. Ensure that the shoots are equally exposed to the light source. Control the temperature and moisture over the course of the experiment.
  • Interpret the results after 2 days. After the shoots have been exposed to the light source for 2 days, interpret the results. The shoots with covered tips will not grow towards the light source, but the other 6 shoots will.
  • Record the amount of growth. To get accurate, quantitative results, you can measure the growth of each shoot and write down the direction of growth.

Phototropism is the growth response of a plant towards or away from light.

There are several experiments that can be done to demonstrate phototropism in plants, including: The experiment with potted plants, where a plant is grown in a pot and then covered on one side with a black paper. The plant will grow towards the light source The experiment with grass seedlings, where grass seedlings are grown in a tray and exposed to light from one side. The seedlings will grow towards the light source The experiment with Avena seedlings, where Avena seedlings are grown in a test tube and exposed to light from one side. The seedlings will bend towards the light source The experiment with coleoptiles, which are the protective sheaths surrounding grass shoots. Coleoptiles are placed in a darkened room and exposed to light from one side. The coleoptiles will bend towards the light source

Phototropism works in plants through the unequal distribution of auxin, a hormone responsible for promoting growth in plants. When light is shone on one side of the plant, it stimulates the cells on that side to produce more auxin. This causes the cells on that side to grow faster, bending the plant in the direction of the light.

The knowledge of phototropism has practical implications for agriculture and horticulture, as it can be used to increase crop yields and improve the growth of ornamental plants. By manipulating the light exposure of plants, farmers and horticulturists can encourage the growth of plants in a desired direction, leading to more efficient use of space and resources.

The study of phototropism is important for future careers in Biology because it provides a fundamental understanding of plant growth and development. This knowledge is essential for careers in areas such as botany, plant sciences, agriculture, horticulture, and other related fields, where an understanding of plant growth is crucial.

The study of phototropism is approached in A-level Biology through a combination of theoretical and practical work. Students learn about the mechanisms of phototropism, the role of hormones in plant growth, and the factors that affect the direction of growth. They also conduct practical experiments to demonstrate phototropism in plants and gain hands-on experience in plant growth and development.

Some of the factors that can affect the direction of phototropism in plants include: The intensity of the light The duration of the light exposure The wavelength of the light The age of the plant The species of the plant

Phototropism can be used to grow plants in space as it provides a way to orient the plants towards a light source, even in a low-gravity environment. This can be important for growing food crops or for conducting experiments on plant growth in space.

Still got a question? Leave a comment

Leave a comment, cancel reply.

Save my name, email, and website in this browser for the next time I comment.

CIE 1 Cell structure

Roles of atp (a-level biology), atp as an energy source (a-level biology), the synthesis and hydrolysis of atp (a-level biology), the structure of atp (a-level biology), magnification and resolution (a-level biology), calculating cell size (a-level biology), studying cells: confocal microscopes (a-level biology), studying cells: electron microscopes (a-level biology), studying cells: light microscopes (a-level biology), life cycle and replication of viruses (a-level biology), cie 10 infectious disease, bacteria, antibiotics, and other medicines (a-level biology), pathogens and infectious diseases (a-level biology), cie 11 immunity, types of immunity and vaccinations (a-level biology), structure and function of antibodies (a-level biology), the adaptive immune response (a-level biology), introduction to the immune system (a-level biology), primary defences against pathogens (a-level biology), cie 12 energy and respiration, anaerobic respiration in mammals, plants and fungi (a-level biology), anaerobic respiration (a-level biology), oxidative phosphorylation and chemiosmosis (a-level biology), oxidative phosphorylation and the electron transport chain (a-level biology), the krebs cycle (a-level biology), the link reaction (a-level biology), the stages and products of glycolysis (a-level biology), glycolysis (a-level biology), the structure of mitochondria (a-level biology), the need for cellular respiration (a-level biology), cie 13 photosynthesis, limiting factors of photosynthesis (a-level biology), cyclic and non-cyclic phosphorylation (a-level biology), the 2 stages of photosynthesis (a-level biology), photosystems and photosynthetic pigments (a-level biology), site of photosynthesis, overview of photosynthesis (a-level biology), cie 14 homeostasis, ectotherms and endotherms (a-level biology), thermoregulation (a-level biology), plant responses to changes in the environment (a-level biology), cie 15 control and co-ordination, the nervous system (a-level biology), sources of atp during contraction (a-level biology), the ultrastructure of the sarcomere during contraction (a-level biology), the role of troponin and tropomyosin (a-level biology), the structure of myofibrils (a-level biology), slow and fast twitch muscles (a-level biology), the structure of mammalian muscles (a-level biology), how muscles allow movement (a-level biology), the neuromuscular junction (a-level biology), features of synapses (a-level biology), cie 16 inherited change, calculating genetic diversity (a-level biology), how meiosis produces variation (a-level biology), cell division by meiosis (a-level biology), importance of meiosis (a-level biology), cie 17 selection and evolution, types of selection (a-level biology), mechanism of natural selection (a-level biology), types of variation (a-level biology), cie 18 biodiversity, classification and conservation, biodiversity and gene technology (a-level biology), factors affecting biodiversity (a-level biology), biodiversity calculations (a-level biology), introducing biodiversity (a-level biology), the three domain system (a-level biology), phylogeny and classification (a-level biology), classifying organisms (a-level biology), cie 19 genetic technology, cie 2 biological molecules, properties of water (a-level biology), structure of water (a-level biology), test for lipids and proteins (a-level biology), tests for carbohydrates (a-level biology), protein structures: globular and fibrous proteins (a-level biology), protein structures: tertiary and quaternary structures (a-level biology), protein structures: primary and secondary structures (a-level biology), protein formation (a-level biology), proteins and amino acids: an introduction (a-level biology), phospholipid bilayer (a-level biology), cie 3 enzymes, enzymes: inhibitors (a-level biology), enzymes: rates of reaction (a-level biology), enzymes: intracellular and extracellular forms (a-level biology), enzymes: mechanism of action (a-level biology), enzymes: key concepts (a-level biology), enzymes: introduction (a-level biology), cie 4 cell membranes and transport, transport across membranes: active transport (a-level biology), investigating transport across membranes (a-level biology), transport across membranes: osmosis (a-level biology), transport across membranes: diffusion (a-level biology), signalling across cell membranes (a-level biology), function of cell membrane (a-level biology), factors affecting cell membrane structure (a-level biology), structure of cell membranes (a-level biology), cie 5 the mitotic cell cycle, chromosome mutations (a-level biology), cell division: checkpoints and mutations (a-level biology), cell division: phases of mitosis (a-level biology), cell division: the cell cycle (a-level biology), cell division: chromosomes (a-level biology), cie 6 nucleic acids and protein synthesis, transfer rna (a-level biology), transcription (a-level biology), messenger rna (a-level biology), introducing the genetic code (a-level biology), genes and protein synthesis (a-level biology), synthesising proteins from dna (a-level biology), structure of rna (a-level biology), dna replication (a-level biology), dna structure and the double helix (a-level biology), polynucleotides (a-level biology), cie 7 transport in plants, translocation and evidence of the mass flow hypothesis (a-level biology), the phloem (a-level biology), importance of and evidence for transpiration (a-level biology), introduction to transpiration (a-level biology), the pathway and movement of water into the roots and xylem (a-level biology), the xylem (a-level biology), cie 8 transport in mammals, controlling heart rate (a-level biology), structure of the heart (a-level biology), transport of carbon dioxide (a-level biology), transport of oxygen (a-level biology), exchange in capillaries (a-level biology), structure and function of blood vessels (a-level biology), cie 9 gas exchange and smoking, lung disease (a-level biology), pulmonary ventilation rate (a-level biology), ventilation (a-level biology), structure of the lungs (a-level biology), general features of exchange surfaces (a-level biology), understanding surface area to volume ratio (a-level biology), the need for exchange surfaces (a-level biology), edexcel a 1: lifestyle, health and risk, phospholipids – introduction (a-level biology), edexcel a 2: genes and health, features of the genetic code (a-level biology), gas exchange in plants (a-level biology), gas exchange in insects (a-level biology), edexcel a 3: voice of the genome, edexcel a 4: biodiversity and natural resources, edexcel a 5: on the wild side, reducing biomass loss (a-level biology), sources of biomass loss (a-level biology), transfer of biomass (a-level biology), measuring biomass (a-level biology), net primary production (a-level biology), gross primary production (a-level biology), trophic levels (a-level biology), edexcel a 6: immunity, infection & forensics, microbial techniques (a-level biology), the innate immune response (a-level biology), edexcel a 7: run for your life, edexcel a 8: grey matter, inhibitory synapses (a-level biology), synaptic transmission (a-level biology), the structure of the synapse (a-level biology), factors affecting the speed of transmission (a-level biology), myelination (a-level biology), the refractory period (a-level biology), all or nothing principle (a-level biology), edexcel b 1: biological molecules, inorganic ions (a-level biology), edexcel b 10: ecosystems, nitrogen cycle: nitrification and denitrification (a-level biology), the phosphorus cycle (a-level biology), nitrogen cycle: fixation and ammonification (a-level biology), introduction to nutrient cycles (a-level biology), edexcel b 2: cells, viruses, reproduction, edexcel b 3: classification & biodiversity, edexcel b 4: exchange and transport, edexcel b 5: energy for biological processes, edexcel b 6: microbiology and pathogens, edexcel b 7: modern genetics, edexcel b 8: origins of genetic variation, edexcel b 9: control systems, ocr 2.1.1 cell structure, structure of prokaryotic cells (a-level biology), eukaryotic cells: comparing plant and animal cells (a-level biology), eukaryotic cells: plant cell organelles (a-level biology), eukaryotic cells: the endoplasmic reticulum (a-level biology), eukaryotic cells: the golgi apparatus and lysosomes (a-level biology), ocr 2.1.2 biological molecules, introduction to eukaryotic cells and organelles (a-level biology), ocr 2.1.3 nucleotides and nucleic acids, ocr 2.1.4 enzymes, ocr 2.1.5 biological membranes, ocr 2.1.6 cell division, diversity & organisation, ocr 3.1.1 exchange surfaces, ocr 3.1.2 transport in animals, ocr 3.1.3 transport in plants, examples of xerophytes (a-level biology), introduction to xerophytes (a-level biology), ocr 4.1.1 communicable diseases, structure of viruses (a-level biology), ocr 4.2.1 biodiversity, ocr 4.2.2 classification and evolution, ocr 5.1.1 communication and homeostasis, the resting potential (a-level biology), ocr 5.1.2 excretion, ocr 5.1.3 neuronal communication, hyperpolarisation and transmission of the action potential (a-level biology), depolarisation and repolarisation in the action potential (a-level biology), ocr 5.1.4 hormonal communication, ocr 5.1.5 plant and animal responses, ocr 5.2.1 photosynthesis, ocr 5.2.2 respiration, ocr 6.1.1 cellular control, ocr 6.1.2 patterns of inheritance, ocr 6.1.3 manipulating genomes, ocr 6.2.1 cloning and biotechnology, ocr 6.3.1 ecosystems, ocr 6.3.2 populations and sustainability, related links.

  • A-Level Biology Past Papers

Boost your A-Level Biology Performance

Get a 9 in A-Level Biology with our Trusted 1-1 Tutors. Enquire now.

100+ Video Tutorials, Flashcards and Weekly Seminars. 100% Money Back Guarantee

Gain hands-on experience of how physics is used in different fields. Boost your university application with our summer programme!

Learn live with other students and gain expert tips and advice to boost your score.

experiment for phototropism

Let's get acquainted ? What is your name?

Nice to meet you, {{name}} what is your preferred e-mail address, nice to meet you, {{name}} what is your preferred phone number, what is your preferred phone number, just to check, what are you interested in, when should we call you.

It would be great to have a 15m chat to discuss a personalised plan and answer any questions

What time works best for you? (UK Time)

Pick a time-slot that works best for you ?

How many hours of 1-1 tutoring are you looking for?

My whatsapp number is..., for our safeguarding policy, please confirm....

Please provide the mobile number of a guardian/parent

Which online course are you interested in?

What is your query, you can apply for a bursary by clicking this link, sure, what is your query, thank you for your response. we will aim to get back to you within 12-24 hours., lock in a 2 hour 1-1 tutoring lesson now.

If you're ready and keen to get started click the button below to book your first 2 hour 1-1 tutoring lesson with us. Connect with a tutor from a university of your choice in minutes. (Use FAST5 to get 5% Off!)

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • v.18(5); 2006 May

Logo of plntcell

Phototropism: Bending towards Enlightenment

Research on phototropism has had far-reaching consequences in the field of plant biology, from helping to refute the ancient misconception of plant insensitivity to the environment to the discovery of the plant hormone auxin and the identification of the phototropin photoreceptors. In this essay, we trace the major trends and ideas that shaped past shoot phototropism research and briefly summarize the current state of the field.

ANCIENT AND MEDIEVAL PERCEPTIONS OF PHOTOTROPISM

For centuries, poets, philosophers, artists, and scientists have noted and studied the phototropic movement of plants. In one of the earliest depictions of plant phototropism, Venus, the ancient goddess of love, transforms Clytie, a water nymph, into a plant because of her infatuation with Apollo, the sun god. Associated with her metamorphosis into a green plant, Clytie turns and follows the movement of Apollo ( Ovid et al., 1998 ). This tale of unrequited love is based on the assumption, developed by the early classical philosophers, that plants exhibit completely passive responses to the environment.

The earliest Greek philosophers, Anaxagoras (500–428 BCE) and Empedocles (495–435 BCE), believed that plants, like animals, are sensitive and capable of motion ( Drossaart Lulofs and Poortman, 1989 ). Although Plato (427–347 BCE) also believed in plant sensitivity, he rejected the idea of plant movement ( Shemp, 1947 ; Plato, 2000 ). Aristotle (384–322 BCE) argued that plants are totally passive and insensitive, and plant insensitivity served as a key criterion for distinguishing between plants and animals ( McKeon, 1947 ; Drossaart Lulofs and Poortman, 1989 ). Following Aristotle's reasoning, Theophrastus (371–287 BCE) also considered plants as passive organisms. In his botanical writings, Theophrastus recorded the phototropic (and solar-tracking) tendencies of plants, but rather than implicating any activators in the plant, he attributed the phenomenon to the sun's activity in removing fluid from the illuminated side of the plant ( Theophrastus, 1976 ). Because Aristotelian scientific philosophy placed greater value on logic alone and downplayed the need for experimental testing, Theophrastus' simple explanation of phototropism persisted until the 17th century when experimental botanists began to recognize plant sensitivity ( Webster, 1966 ).

During the middle ages, herbalists were more interested in the medicinal properties of plants than understanding plant biology. According to the doctrine of signatures, which associated the shape of a plant with its medicinal usage, phototropic plants may have been prescribed for the treatment of snake and serpent bites due to the serpentine shapes they display ( von Erhardt-Siebold, 1937 ). From today's scientific standpoint, the most significant advancement of the medieval herbals was the establishment of a nomenclature that separated plants whose flower-opening is dependent on the sun (composite) from plants that display shoot and leaf phototropism (solago) ( von Erhardt-Siebold, 1937 ).

DISCOVERING THE INDUCTIVE NATURE OF PHOTOTROPISM

During the renaissance, some early scientists began studying “natural magic,” which was reliant on the elements and occult properties of material things. In contrast with the Aristotelian disdain of experimentation, these early scientists used experimental observation in addition to classical texts to guide their thinking. Giambattista della Porta (1535–1615), probably one of the most well-known practitioners of natural magic, experimented with movement responses of cucumber seedlings. Drawing on Theophrastus' description of phototropism and anthropomorphic treatment of the response by medieval sources, della Porta described plant phototropism as a “rejoicing” response to the sun ( della Porta, 1569 ). Furthermore, in an attempt to explain seemingly similar natural phenomenon, he proposed that the same fundamental law of nature, which he called “sympathy,” governed the attraction of iron toward magnets, hens toward eggs, and the phototropic movement of plants toward the sun ( della Porta, 1569 ). Although it is unclear if della Porta actually believed the concept of plant sensitivity, his explanation of phototropism as a rejoicing and sympathetic response helped open the debate on plant sensitivity.

Francis Bacon (1561–1626), who helped shape the modern scientific method, was familiar with della Porta's writing about plant movement. Bacon recorded the tropistic movements of many different plants but held to the classical belief in plant insensitivity. As such, Bacon discarded della Porta's explanation of plant phototropism as a sympathetic or rejoicing response to the sun, and, like Theophrastus, he viewed phototropism as a simple mechanical consequence of wilting. He wrote, “the cause (of phototropism) is somewhat obscure…the part beateth by the sun waxeth more faint and flaccid in the stalk, and less able to support the flower” ( Bacon et al., 1627 ).

Intrigued by Bacon's discussion of plants, Thomas Browne (1605–1685) began studying plant physiology. As an alchemist, Browne was seeking a mystical unification of the universe and was more open to the idea of plant sensitivity than Bacon ( Webster, 1966 ). At the time, plants were thought to emit “bad air,” so Browne believed that plant movements allowed plants to avoid the bad air produced by neighboring plants. In what is probably the first crude scientific experiment on plant phototropism, Brown observed that mustard seedlings grown in front of a basement window would eventually reorient themselves toward the window after he rotated the pot ( Browne, 1658 ). When Robert Sharrock (1630–1684) repeated Browne's experiment, he concluded that the response was stimulated by fresh air and caused by growth rather than a mechanical consequence of wilting ( Sharrock, 1672 ). These early phototropism experiments are significant because they provided some of the first scientific examples of plant sensitivity.

As the leading botanical taxonomist, John Ray (1628–1705) would have recognized the taxonomic implications that plant sensitivity would have on the classical distinction between plants and animals. To escape the problem this created, Ray considered phototropic movement of plants to be a mechanical effect of temperature on growth. Believing that the rapid etiolated growth of dark-grown seedlings was caused by warmer temperatures, Ray argued that phototropism is caused by a temperature gradient across the seedling with the side closest to the window being colder and slower-growing ( Ray et al., 1686 ; Sachs et al., 1890 ). Although Ray's temperature hypothesis was later disregarded, his proposition that etiolated growth and phototropism are somehow connected remains a topic of investigation.

Due to the previous work of Browne and the discovery of the sensitive Mimosa plant, the idea of plant sensitivity to light began to receive wider acceptance ( Webster, 1966 ). However, responsiveness to light was still considered to be mechanical rather than inductive. Although Charles Bonnet (1720–1793) attributed the process of photomorphogenesis to light, he failed to recognize the importance of light for phototropism. Instead, his phototropism experiments ( Figure 1 ) led him to believe that plants were turning toward the warmth of the sun ( Bonnet, 1754 , 1779 ; Sachs et al., 1890 ). Yet based on Bonnet's observations, Henri-Louis DuHamel (1700–1782) concluded that light rather than warmth is more important for the response ( DuHamel, 1758 ; Sachs et al., 1890 ).

An external file that holds a picture, illustration, etc.
Object name is tpc1801110f01.jpg

Woodcut Depicting a Charles Bonnet Tropism Experiment from 1779.

Two etiolated bean seedlings ( a and b ) oppositely placed in a vase ( v ) of water were tied downward ( e ). With the shutter ( f ) closed, each seedling reoriented upward toward the nearest wall (seedling a toward wall q and seedling b toward wall p ). When the shutter was raised, both seedlings reoriented toward the opening ( o ). Reprinted from Bonnet (1779) .

The Romantic period of the late 18th and early 19th centuries was characterized by a philosophical backlash against the mechanistic view of life that had dominated the enlightenment, and the concept of an endogenous “vital force” served as a common explanation of plant phenomena ( Sachs et al., 1890 ). As this outlook began losing popularity between 1820 and 1840, Henri Dutrochet (1776–1843) proposed that phototropism was an inductive response to light ( Dutrochet, 1824 , 1826 , 1828 , 1837 ). However, Dutrochet's contemporary, Augustin Pyramus de Candolle (1778–1841), thought that phototropism was simply a mechanical consequence of greater etiolated growth on the shaded side of the plant ( de Candolle, 1832 ). This explanation of phototropism was challenged by Albert Bernard Frank (1839–1900), who was the first to propose that phototropism and gravitropism are inductive responses sharing a common underlying process ( Frank, 1868 ). The inductive nature of the response was finally confirmed when Julius von Wiesner (1838–1919) showed that plants continue to bend toward a light source even after the light is turned off ( von Wiesner, 1878 ).

Charles Darwin (1809–1882) further explored the inductive nature and mechanistic connection between phototropism and gravitropism. He proposed that the back and forth circumnutation associated with plant growth could be directed by a stimulus such as light or gravity ( Darwin, 1880 ). Although Darwin's circumnutation theory of tropism served to propose a common mechanism underlying gravitropism and phototropism, the most significant discovery from his studies of plant movements was his demonstration that the site of photoperception at the shoot tip and the location of curvature are separable. From his observations, Darwin was able to propose that a transmissible substance produced in the tip is responsible for inducing curvature in lower regions of the plant ( Darwin, 1880 ). This insightful discovery eventually lead to the discovery of the first plant hormone, auxin.

THE DISCOVERY OF AUXIN AND UNDERSTANDING ITS ROLE IN PHOTOTROPISM

Darwin's ideas were initially dismissed by other plant physiologists (reviewed in Heslop-Harrison, 1980 ). Nevertheless, evidence in favor of Darwin's transmissible substance began to accumulate when Rothert (1894) also showed that light sensitivity is greatest near the tip of maize coleoptiles. Subsequent results of Fitting (1907) , Boysen-Jensen (1911) , and Paal (1918) provided more direct evidence that a transmissible substance produced in the tip participates in the response. This research culminated in a model put forth independently by Cholodny (1927) and Went (1926 , 1928 ), which proposed that light-mediated lateral redistribution of a plant growth hormone to the shaded side of the seedling causes the differential growth associated with phototropic curvature. This growth substance was shortly identified from human urine by Kogl and Haagen-Smit (1931) , who named the hormone auxin, derived from the Greek verb auxein, meaning “to grow.”

Although the Cholodny-Went model has remained the dominant explanation of phototropism, other models have challenged its validity. Historically, the model advocated by A.H. Blaauw has been one of the most common alternatives to the Cholodony-Went theory. Similar to Candolle, Blaauw proposed that phototropism is a secondary consequence of differential growth inhibition associated with photomorphogenesis ( Blaauw, 1919 ). While arguing against the Blaauw theory as the sole basis for phototropism, Boysen-Jensen et al. (1936) and Went and Thimann (1937) held out the possibility that both models might function during the phototropic response of dicotyledonous plants. Evidence for this complementary Blaauw/Cholodony-Went model was based on the work of Overbeek, who demonstrated that although unilateral light stimulates the movement of auxin to the shaded side of the hypocotyl, up to half the differential growth associated with phototropism can be attributed to light-mediated growth inhibition ( Overbeek, 1932 , 1933 ). More detailed measurements of growth during the 1980s supported this view by indicating that growth inhibition on the illuminated side is accompanied by growth promotion on the shaded side of seedlings ( Iino and Briggs, 1984 ; Macleod et al., 1985 ; Rich et al., 1985 ; Baskin, 1986 ). However, Cosgrove (1985) demonstrated that light-mediated growth inhibition occurs much sooner than the onset of phototropic curvature, a finding that would not be predicted by the Blaauw model. Additional evidence against the Blaauw model came from a study by Liscum et al. (1992) , which demonstrated that growth inhibition and phototropism are separable.

Another major challenge to the Cholodny-Went model came from an experiment indicating that carotenoids (proposed phototropism photoreceptors) participate in photoinactivation of auxin ( Kogl and Schuringa, 1944 ). This observation led to the hypothesis that phototropism is caused by differential carotenoid-mediated auxin inactivation (reviewed in Shank, 1950 ). Evidence against the auxin inactivation model came when Briggs et al. (1957) failed to see any in vivo change in total auxin concentration following treatment with light. Instead, their results showed that a barrier inserted between the illuminated and dark sides of the coleoptile prevents the development of an auxin gradient under unilateral illumination. Further support of the Cholodony-Went model came when Briggs (1963a) reported a correlation between an auxin gradient and the magnitude of phototropic response and when Pickard and Thimann (1964) showed that unilateral blue light causes greater radiolabeled auxin accumulation on the shaded flank of coleoptiles. Although these key studies demonstrated that a differential auxin gradient is associated with phototropism, they could not resolve how this gradient develops.

According to the classical Cholodny-Went model, lateral auxin transport gives rise to phototropism. However, experiments by Shen-Miller and Gordon ( Shen-Miller and Gordon, 1966 ; Shen-Miller et al., 1969 ) indicated that light inhibits polar auxin transport, which led them to propose that phototropism is caused by light-mediated inhibition of polar auxin transport on the illuminated flank of a seedling. Related to this model, Naqvi (1972) proposed that light-induced production of abscisic acid on the illuminated flank of the seedling causes unequal polar auxin transport. To test the idea that unequal polar auxin transport gives rise to phototropism, Gardner et al. (1974) followed radiolabeled auxin that had been asymmetrically applied to coleoptiles. Although they were not able to completely exclude a role for light-mediated inhibition of polar auxin transport, their results further demonstrated that unilateral blue light induced lateral auxin transport. At the same time, Kang and Burg (1974) reported that the enhancement of pea epicotyl phototropism by red light or gibberellins did not correlate with an increase in lateral auxin transport and proposed that the magnitude of the phototropic response is determined by adjusting auxin sensitivity. Altogether, these studies supported the involvement of auxin in phototropism, but the precise mechanism of how auxin transport and signaling cause phototropism remained unknown.

Many more details about how changes in auxin transport influence phototropism are now emerging from research using Arabidopsis as a model system. One report claims that mutations in PIN3 , which encodes an auxin efflux carrier involved in lateral auxin transport, can disrupt phototropism ( Friml et al., 2002 ). However the relationship between blue light signaling and PIN3 is uncertain. Work by Blakeslee et al. (2004) suggests that blue light does not effect the localization of PIN3 as it does for PIN1, an auxin efflux carrier thought to be important for polar auxin transport. Light-mediated relocalization of PIN1 appears to play a role in phototropism, as mutants in MDR1 , a gene encoding a P-glycoprotein ABC transporter, show less PIN1 localization to the basal end of hypocotyl cells and an enhanced phototropic response ( Noh et al., 2003 ). With less PIN1 localized to the basal end of hypocotyl cells, it was proposed that by decreasing polar auxin transport, phototropism may be enhanced by increasing the amount of auxin available for lateral transport ( Noh et al., 2003 ). However, some polar auxin transport appears to be important for the normal progression of phototropism as big mutants, which have diminished polar auxin transport, display longer phototropic latent periods ( Whippo and Hangarter, 2005 ). Clearly, more research is needed to determine how changes in both lateral and polar auxin transport impact phototropism. In particular, it will be important to understand how the dynamics of auxin transport are regulated in the context of the differential growth responses that lead to curvature.

Although our knowledge about auxin transport has advanced significantly, we have an even better understanding of auxin signaling during phototropism and a compelling model is developing. An important study by Harper et al. (2000) showed that the nonphototropic hypocotyl4 ( nph4 ) locus encodes Auxin Response Factor7 (ARF7), a member of the auxin response factor family. ARFs function as transcriptional regulators whose activity is inhibited by binding to AUX/IAA proteins. Auxin facilitates ARF activity by promoting the targeting of AUX/IAA proteins for degradation via the ubiquitin-proteosome pathway ( Liscum and Reed, 2002 ). With respect to phototropism, the degradation of IAA19, which binds to and inactivates ARF7, participates in phototropism under low light conditions ( Tatematsu et al., 2004 ). However, ARF7 involvement in phototropism does not seem absolutely necessary, since nph4/arf7 mutants display a phototropic response when treated with ethylene or given a red light pretreatment ( Stowe-Evans et al., 2001 ). This suggests that red light or ethylene activates another ARF that functions during phototropism ( Harper et al., 2000 ). A possible scenario is that the ARFs promote phototropism by controlling the expression of genes containing auxin response elements (AuxREs) ( Liscum and Reed, 2002 ). In support of this scenario, a recent study found that eight genes with differential transcript accumulation across phototropic-stimulated Brassica oleracea hypocotyls have one or more AuxREs ( Esmon et al., 2005 ). Interestingly, two of these genes encode expansins, which are involved in cell wall extension ( Esmon et al., 2005 ). Therefore, this study provides a plausible mechanism linking the differential growth underlying phototropism to the auxin regulation of expansin activity ( Esmon et al., 2005 ). Since the genes encoding for AUX/IAA proteins also contain this AuxRE, a negative feedback loop involving the upregulation of AUX/IAA proteins may participate in the progression of normal phototropism ( Tatematsu et al., 2004 ). For example, the reversal in the differential growth gradient that prevents the hypocotyl from curling around upon itself as the position of curvature migrates down the length of the hypocotyl ( Silk, 1984 ; Whippo and Hangarter, 2003 ) may be a partial consequence of ARF-mediated upregulation of AUX/IAA proteins.

THE SEARCH FOR A PHOTOTROPISM PHOTORECEPTOR

In parallel with research on the role of auxin in phototropism, another important area concerns how plants perceive a unilateral light source. As soon as it became more widely accepted that phototropism is stimulated by light in the 1800s, the focus turned toward identifying the property of light responsible. As early as 1817, Sebastiano Poggioli reported that blue wavelengths of light are more effective at orienting plant growth ( Poggioli, 1817 ). After several conflicting studies by Payer (1842) , Zantedeschi (1843) , Guillemin (1858) , and Sachs (1864) , Julius von Wiesner published the first methodical examination of the phototropism action spectra at specific wavelengths of light ( Figure 2 ) and confirmed that phototropism is strongest toward blue/violet light ( von Wiesner, 1878 ). Blaauw (1909) later identified a peak wavelength around 450 nm. Johnston elaborated on Blaauw's action spectra and identified dual peaks at 480 and 440 nm ( Johnston, 1934 ).

An external file that holds a picture, illustration, etc.
Object name is tpc1801110f02.jpg

Julius von Wiesner's 1878 Diagram of the Action Spectra for Phototropism.

von Wiesner observed the phototropic response of pea (Wicke), cress (Kresse), and grass (Weide) seedlings to various light qualities. Letters along the x axis represent Fraunhofer lines demarking light quality between 759 nm (A) and 396 nm (H). Reprinted from von Wiesner (1878) .

When the action spectra for phototropism became better defined, attention turned toward identification of the blue light photoreceptor responsible for the response. Because the phototropism action spectra resemble the absorption spectra of carotenoids ( Haig, 1935 ; Wald and Du Buy, 1936 ) and carotenoid concentration is greatest in the tips where phototropism sensitivity is greatest, carotenoids were originally considered to be the pigments responsible for phototropism (reviewed in Shank, 1950 ). This hypothesis went relatively unchallenged until Galston and Baker (1949) proposed the involvement of a flavin or flavoprotein photoreceptor during phototropism. The flavin hypothesis was primarily based on observations that flavins have a peak absorbance around 450 and that flavonoids can inactivate auxin in vitro ( Galston, 1950 ).

For nearly the next 50 years, scientists would continue to debate the identity of the phototropism photoreceptor without convincing proof for or against either a flavonoid or carotenoid photoreceptor. One unfruitful line of research proposed that the blue light photoreceptor would be uncovered via studies of light-induced absorbance changes ( Berns and Vaughn, 1970 ; Munoz and Bulter, 1975 ; Brain et al., 1977 ). Another line of research used chemical inhibitors to try to elucidate the identity of the blue light photoreceptor. Schmidt et al. (1977) and Vierstra and Poff (1981a) showed that maize coleoptiles treated with flavin inhibitors fail to display a phototropic response. However, carotenoids also appeared to play a role, perhaps as screening pigments, because coleoptiles treated with the carotenoid biosynthesis inhibitor norflurazon also show a reduced phototropic response ( Vierstra and Poff, 1981b ). In another study, Leong and Briggs (1982) hypothesized the involvement of a flavin cytochrome complex in the perception of light for phototropism because treatment with the electron transport chain inhibitor acriflourfen increased the sensitivity to unilateral blue light.

Eventually, Briggs and associates turned to a biochemical approach in an attempt to identify the blue light photoreceptor. Their work led to the identification of a 120-kD membrane-bound protein whose phosphorylation state and activation is altered by blue light in a fashion that correlated with phototropism ( Gallanger et al., 1988 ; Short and Briggs, 1990 ; Reymond et al., 1992a ; Short et al., 1992 ). Stronger evidence that this blue light phosporylated protein is involved in phototropism came when Reymond et al. (1992b) reported that microsomes from the phototropism mutant JK224 contain lower levels of this 120-kD protein. The JK224 mutant was originally isolated by Khurana and Poff (1989) , who showed that it had a higher threshold requirement for the induction of phototropism and hypothesized that the mutant gene represented a photoreceptor involved in phototropism. Liscum and Briggs (1995) later identified other alleles of JK224 in their screen for nph mutants, strengthening this hypothesis.

Cloning of nph1 confirmed the prediction that this locus encodes the 120-kD protein ( Huala et al., 1997 ), and subsequent biochemical evidence indicated that two domains of this protein, LOV1 and LOV2, bind to flavin chromophores with spectral properties consistent with phototropism ( Huala et al., 1997 ; Christie et al., 1998 , 1999 ). Hence, this protein was renamed phototropin ( Christie et al., 1999 ), and the carotenoid-based photoreceptor hypothesis lost considerable ground. Still another photoreceptor was predicted to be involved in the induction of high-light phototropism because etiolated phot1 mutants retain a strong phototropic response to long-term irradiation with high-intensity blue light ( Sakai et al., 2000 ). A subsequent study showed that another member of the phototropin family, phot2, functions redundantly to phot1 in the induction of high-light phototropism ( Sakai et al., 2001 ).

With the identification of the phototropins as the phototropism photoreceptors, focus has turned to understanding their mechanism of light perception. Localization experiments reveled that more phot1 is located near the tip of etiolated seedlings than basally ( Sakamoto and Briggs, 2002 ; Knieb et al., 2004 ), thus providing a possible reason for why Darwin observed greater sensitivity toward the tip ( Darwin, 1880 ; Knieb et al., 2004 ). Salomon et al. (2000) presented evidence that light causes the formation of an adduct between a Cys residue located in the LOV domain and the flavin chromophore. In terms of the induction of phototropism and the light-mediated autophosphorylation of phot1, the LOV2 domain appears more critical than the LOV1 domain ( Christie et al., 2002 ). However, how light-mediated autophosporylation leads to a phototropic response remains to be seen. One possibility is that a gradient of autophosphorylated phototropin across the seedling precedes the development of phototropic curvature ( Salomon et al., 1997 ).

At this point, we know very little about the signaling components immediately downstream of the phototropins. NPH3 and RPT2, two related proteins with unknown function, bind to phot1 ( Motchoulski and Liscum, 1999 ; Inada et al., 2004 ) and function very early in phototropin signaling. These proteins are clearly important for phototropism since of all the characterized phototropism mutants, only nph3 mutants fail to show a response under any light condition ( Liscum and Briggs, 1996 ; Sakai et al., 2000 ). While RPT2 is not necessary for phototropism, it participates in the promotion of high-light phototropism ( Sakai et al., 2000 ). Future studies of NPH3, RPT2, and possibly other members of this family of proteins may become critical in uncovering how phototropin signaling modulates auxin transport. Calcium signaling has also been implicated as an early component of phototropin signaling ( Gehring et al., 1990 ; Babourina et al., 2004 ). Although phot1 is required for a rapid blue light–mediated increase in cytosolic calcium ( Baum et al., 1999 ) cytosolic calcium may not be necessary for phototropism ( Folta et al., 2003 ). Instead, the phot1-induced increase in calcium seems to be more directly related to rapid phot1-mediated growth inhibition in hypocotyls ( Folta et al., 2003 ).

PHYTOCHROME AND CRYPTOCHROME SIGNALING IN THE PROMOTION OF PHOTOTROPISM

The phototropins are not the only photoreceptors involved in phototropism. Although red light does not typically induce phototropism, a series of studies by Curry (1957) , Blaauw-Jansen (1959) , Asomaning and Galston (1961) , and Briggs (1963b) showed that pretreating seedlings with red light modulates phototropic sensitivity to unilateral blue light. Implicating the phytochrome red/far-red reversible photoreceptors in phototropism, Briggs (1963b) found that the red light enhancement of phototropism can be reversed by far-red light. This was later supported by a spectral correlation between phytochrome and the red light modification of phototropism ( Chon and Briggs, 1966 ). Although these studies implicated a phytochrome role in phototropism, they could not address the relative importance of phytochrome activity in the promotion of phototropism.

Several studies indicated that phytochromes can play more than just a secondary role in phototropism under some circumstances. For example, Iino et al. (1984) proposed that phytochromes may be essential for phototropism in some cases because they observed a phototropic response of maize mesocotyls to unilateral red light. Since phytochromes also absorb blue light and saturating red or far-red light from above could inhibit the phototropic response of pea epicotyls, Parker et al. (1989) also concluded that pea epicotyl phototropism is primarily induced by phytochromes in conjunction with an unknown blue light photoreceptor, then referred to as cryptochrome, playing a secondary role. However, red light is generally much less effective at inducing phototropism ( von Wiesner, 1878 ; Blaauw, 1909 ), so the phytochromes are not thought to be the directional photoreceptors in phototropism.

Several studies using Arabidopsis confirmed a significant role for the phytochromes, not only in the red light enhancement of phototropism ( Parks et al., 1996 ; Hangarter, 1997 ; Janoudi et al., 1997a , 1997b ; Stowe-Evans et al., 2001 ) but also in the absence of a red light pretreatment ( Hangarter, 1997 ; Whippo and Hangarter, 2004 ). Likewise, other studies indicate that the cryptochrome blue light photoreceptors can enhance the development of phototropic curvature ( Ahmad et al., 1998 ; Lasceve et al., 1999 ; Whippo and Hangarter, 2003 ). Since the phytochromes and cryptochromes function in parallel and somewhat redundantly during other light-mediated responses, they may be affecting phototropism in a similar fashion. Indeed, phytochrome cryptochrome double mutants display a severely reduced phototropic response under light conditions where phytochrome and cryptochrome single mutants show normal responses (C.W. Whippo and R.P. Hangarter, unpublished results). It is possible that the regulation of HY5, a transcriptional activator mainly associated with photomorphogenesis, by the phytochromes and cryptochromes participates in the promotion of phototropism because hy5 mutants display a significantly slower phototropic response to very low light treatments ( Whippo and Hangarter, 2005 ).

TOWARD UNDERSTANDING PHOTOTROPISM SENSITIVITY AND RESPONSIVENESS

The focus of phototropism research over the last 150 years was primarily concerned with the mechanistic aspects of the response. However, the degree to which a plant or plant part responds to unilateral light can vary widely. In some cases, different phototropic responses are a trivial result of mechanics: a large diameter shoot requires more differential growth than a small diameter shoot to reach the same angle of curvature. In other cases, differences in phototropism are more connected to the molecular physiology associated with changes in sensitivity and acclimation to prevailing light. Regulation of phototropism sensitivity/responsiveness can manifest itself in several different ways under long-term exposure to light or brief pulses of light.

The observation that etiolated seedlings exposed to continuous unilateral bright light have a slower response than seedlings exposed to continuous dim light was first observed by von Wiesner (1878) . As already discussed, von Wiesner was investigating whether phototropism is a passive mechanical consequence of light or an inductive response. According to the hypothesis that phototropism is a passive mechanical phenomenon, brighter light was expected to cause a faster, stronger response. When testing this predication of the mechanical phototropism model, von Wiesner (1878) observed that increasing the light intensity past a certain threshold actually retarded the phototropic response of etiolated seedlings. However, von Wiesner also believed that the induction of phototropism was closely related to the induction of photomorphogenesis and concluded that the regulation of phototropic sensitivity and responsiveness coincides with light-mediated growth inhibition ( von Wiesner, 1878 ). During the next century, Pringsheim (1912) and Ellis (1987) reconfirmed von Wiesner's initial observation, but a clear explanation of the phenomenon was not evident.

In a series of articles, Whippo and Hangarter (2003 , 2004 , 2005 ) studied the attenuating effect of continuous high light on the phototropic response of etiolated hypocotyls. From the results of these studies, it appears that there are at least two light-signaling pathways contributing to the attenuation of high-light phototropism. First, high light was found to cause a rapid phototropin-mediated decrease in phototropic responsiveness, as phot1 mutants display a fairly rapid high-light response ( Whippo and Hangarter, 2003 ). This may be due to high light–mediated desensitization of the phototropin photoreceptors or downstream effectors. Interestingly, RPT2 appears to partially mitigate the phot1-mediated attenuation of phototropism because RPT2 is not required for a high-light response in the absence of PHOT1 ( Sakai et al., 2000 ). Secondly, after the rapid phot1-mediated attenuation, the cryptochromes and phytochrome A were found to help maintain a slower high-light phototropic response ( Whippo and Hangarter, 2003 , 2004 ). The cryptochromes and phytochrome A probably function somewhat redundantly in this process by regulating the photomorphogenic regulation of auxin signaling and transport ( Whippo and Hangarter, 2005 ).

Although signaling elements associated with photomorphogenesis participate in the attenuation of high-light phototropism in etiolated seedlings, the relationship between development and phototropism is complex. In contrast with etiolated seedlings, Pringsheim (1912) observed that light-grown pea seedlings are more responsive to brighter light than to dimmer light. Similar behavior was observed in buckwheat ( Ellis, 1987 ) and Arabidopsis ( Whippo and Hangarter, 2005 ). However, how this developmental shift in phototropic sensitivity occurs is currently unknown.

Additional discoveries during the first decades of the 20th century also demonstrated that phototropic responsiveness to brief pulses of light is not as straightforward as might be expected. Using light pulses of varying duration or intensity, Blaauw (1909) determined that the magnitude of the response can be proportional to light dosage. Yet, Pringsheim (1909) and Clark (1913) soon reported that the reciprocity described by Blaauw is only valid under fairly low dosages of light (first positive phototropism). When further increasing the light dosage, Pringsheim (1909) and Clark (1913) stimulated a weak negative response (first negative phototropism), while even higher amounts of light restored a positive phototropic response (second positive phototropism).

Following the adoption of Arabidopsis as a model organism for plant biology research during the 1980s, Poff and associates conducted a series of detailed studies characterizing how light wavelength and dosage affect the first and second positive phototropic responses of etiolated Arabidopsis hypocotyls ( Steinitz and Poff, 1986 ; Konjevic et al., 1989 ; Janoudi and Poff, 1990 , 1991 , 1992 , 1993 ; Janoudi et al., 1992 ). These important physiological studies demonstrated that light signaling involved in pulse-induced phototropism is complex, with both additive and opposing effects on the development of hypocotyl curvature ( Janoudi and Poff, 1993 ). Salomon et al. (1997) have proposed that the complex fluence response curve associated with pulse-induced phototropism is related to the localization and magnitude of phototropin phosphorylation. However, this model awaits testing, and more research is needed to understand the underlying basis for the complex fluence curve of pulse-induced phototropism.

CONCLUSIONS

The history of phototropism is long and rich. Our current understanding of the response has its roots in ancient Greek philosophy and stems from the early physiological studies of the enlightenment. Recent research with Arabidopsis has tremendously expanded our mechanistic understanding of phototropism. We can no longer view the response as a simple or linear physiological response. Instead, phototropism must be viewed as a complex biological response involving interactions of multiple photoreceptors, multiple hormones, and multiple signaling pathways that together orchestrate the establishment of coordinated differential growth gradients. Given its complexity, much phototropism research remains to be done before we can understand all of the underlying mechanisms and know the full account of its biological significance.

Acknowledgments

We thank Nancy Eckardt for her excellent and thoughtful editing. We are supported by grants from the National Science Foundation (IBN-0080783), the Department of Energy (DE-FG02-01ER15223) (R.P.H.), and the Indiana University Briggs Developmental Biology Fellowship (C.W.W.).

  • Ahmad, M., Jarillo, J.A., Smirnova, O., and Cashmore, A.R. (1998). Cryptochrome blue-light photoreceptors of Arabidopsis implicated in phototropism. Nature 392 720–723. [ PubMed ] [ Google Scholar ]
  • Asomaning, E.J.A., and Galston, A.W. (1961). Comparative study of phototropic response and pigment content in oat and barley coleoptiles. Plant Physiol. 36 453–464. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Babourina, O., Godfrey, L., and Voltchanskii, K. (2004). Changes in ion fluxes during phototropic bending of etiolated oat coleoptiles. Ann. Bot. (Lond.) 94 187–194. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bacon, F., Rawley, W., and Cecil, T. (1627). Sylva sylvarum: Or A Natural History in Ten Centuries. (London: John Haviland and Augustine Mathewes).
  • Baskin, T.I. (1986). Redistribution of growth during phototropism and nutation in the pea epicotyl. Planta 169 406–414. [ PubMed ] [ Google Scholar ]
  • Baum, G., Long, J.C., Jenkins, G.I., and Trewavas, A.J. (1999). Stimulation of the blue light phototropic receptor NPH1 causes a transient increase in cytosolic Ca 2+ . Proc. Natl. Acad. Sci. USA 96 13554–13559. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Berns, D.S., and Vaughn, J.R. (1970). Studies of the photopigment system in Physomyes. Biochem. Biophys. Res. Commun. 39 1094–1103. [ PubMed ] [ Google Scholar ]
  • Blaauw, A.H. (1909). Die perzeption des lichtes. Recueil Trav. Bot. Neerland. 5 209–372. [ Google Scholar ]
  • Blaauw, A.H. (1919). Licht und wachstum III. Meded. Landbouwhogeschool Wageningen 15 89–204. [ Google Scholar ]
  • Blaauw-Jansen, G. (1959). The influence of red and far red light on growth and phototropism of the avena seedling. Acta Bot. Neerland . 8 1–39. [ Google Scholar ]
  • Blakeslee, J.J., Bandyopadhyay, A., Peer, W.A., Makam, S.N., and Murphy, A.S. (2004). Relocalization of the PIN1 auxin efflux facilitator plays a role in phototropic responses. Plant Physiol. 134 28–31. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bonnet, C. (1754). Recherches sur L'Usage des Feuilles dans les Plantes et sur Quelques Autres Sujets Relatifs à l'Histoire de la Vegetation. (Gottingue; Leide: E. Luzac).
  • Bonnet, C. (1779). Œuvres d'Histoire Naturelle et de Philosophie. (Neuchatel, Switzerland: S. Fauche).
  • Boysen-Jensen, P. (1911). La transmission de l'irritation phototropique dans l'avena. Bull. Acad. R. Sci. Lett. 3 1–24. [ Google Scholar ]
  • Boysen-Jensen, P., Avery, G.S., Burkholder, P.R., Creighton, H.B., and Scheer, B.A. (1936). Growth Hormones in Plants, 1st ed. (New York: McGraw-Hill Book Company).
  • Brain, R.D., Freeberg, J.A., Weiss, C.V., and Briggs, W.R. (1977). Blue light-induced absorbance changes in membrane fractions from corn and Neurospora. Plant Physiol. 59 948–952. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Briggs, W.R. (1963. a). Mediation of phototropic responses of corn coleoptiles by lateral transport of auxin. Plant Physiol. 38 237–247. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Briggs, W.R. (1963. b). Red light, auxin relationships, and the phototropic responses of corn and oat coleoptiles. Am. J. Bot. 50 196–207. [ Google Scholar ]
  • Briggs, W.R., Tocher, R.D., and Wilson, J.F. (1957). Phototropic auxin redistribution in corn coleoptiles. Science 126 210–212. [ PubMed ] [ Google Scholar ]
  • Browne, T. (1658). The Garden of Cyrus. (London: Golden Cockerel Press).
  • Cholodny, N. (1927). Wuchshormone und tropismem bei den planzen. Biol. Zentralbl. 47 604–626. [ Google Scholar ]
  • Chon, H.P., and Briggs, W.R. (1966). Effect of red light on the phototropic sensitivity of corn coleoptiles. Plant Physiol. 41 1715–1724. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Christie, J.M., Reymond, P., Powell, G.K., Bernasconi, P., Raibekas, A.A., Liscum, E., and Briggs, W.R. (1998). Arabidopsis NPH1: A flavoprotein with the properties of a photoreceptor for phototropism. Science 282 1698–1701. [ PubMed ] [ Google Scholar ]
  • Christie, J.M., Salomon, M., Nozue, K., Wada, M., and Briggs, W.R. (1999). LOV (light, oxygen, or voltage) domains of the blue-light photoreceptor phototropin (nph1): Binding sites for the chromophore flavin mononucleotide. Proc. Natl. Acad. Sci. USA 96 8779–8783. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Christie, J.M., Swartz, T.E., Bogomolni, R.A., and Briggs, W.R. (2002). Phototropin LOV domains exhibit distinct roles in regulating photoreceptor function. Plant J. 32 205–219. [ PubMed ] [ Google Scholar ]
  • Clark, O.L. (1913). Uber negativen phototropismus bei Avena sativa. Zeitschr. Bot. 5 737–770. [ Google Scholar ]
  • Cosgrove, D.J. (1985). Kinetic separation of phototropism from blue-light inhibition of stem elongation. Photochem. Photobiol. 42 745–751. [ PubMed ] [ Google Scholar ]
  • Curry, G.M. (1957). Studies of the Spectral Sensitivity of Phototropism. PhD dissertation (Cambridge, MA: Harvard University).
  • Darwin, C. (1880). The Power of Movement in Plants. (London: John Murray Publishers).
  • de Candolle, A.P. (1832). Physiologie Végétale. (Paris: Béchet).
  • della Porta, G. (1569). Magiæ Naturalis. (London: Rouille).
  • Drossaart Lulofs, H.J., and Poortman, E.L.J., eds. (1989). De Plantis: Five Translations/Nicolaus Damascenus. (New York: North-Holland).
  • DuHamel, H.L. (1758). La Physique des Arbres. (Paris: Chez H.L. Guerin & L.F. Delatour).
  • Dutrochet, H. (1824). Recherches Anatomiques et Physiologiques sur la Structure Intime des Animaux et des Végétaux et sur Leur Motilité. (Paris: J.-B. Baillière). [ PMC free article ] [ PubMed ]
  • Dutrochet, H. (1826). L'Agent Immédiat du Mouvement Vital Dévoilé dans Sa Nature et dans Son Mode d'Action, Chez les Vegétaux et Chez les Animaux. (Paris: J.-B. Bailliere).
  • Dutrochet, H. (1828). Nouvelles Recherches sur l'Endosmose et l'Exosmose Suivies de l'Application Expérimentale de ces Actions Physiques a la Solution du Problême de l'Irritabilité Végétale et a la Détermination de la Cause de l'Ascension des Tiges et de la Descente des Racines. (Paris: J.-B. Bailliere).
  • Dutrochet, H. (1837). Mémoires pour Servir à l'Histoire Anatomique et Physiologique des Végétaux et des Animaux. (Paris: J.-B. Baillière).
  • Ellis, R.J. (1987). Comparison of fluence-response relationships of phototropism in light- and dark-grown buckwheat. Plant Physiol. 85 689–692. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Esmon, C.A., Tinsley, A.G., Ljung, K., Sandberg, G., Hearne, L.B., and Liscum, E. (2005). A gradient of auxin and auxin-dependent transcription precedes tropic growth responses. Proc. Natl. Acad. Sci. USA 103 236–241. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Fitting, H. (1907). Die leitung tropisticher reize in parallelotropen planzenteilen. Jahr. Wiss. Bot. 44 177–253. [ Google Scholar ]
  • Folta, K.M., Lieg, E.J., Durham, T., and Spalding, E.P. (2003). Primary inhibition of hypocotyl growth and phototropism depend differently on phototropin-mediated increases in cytoplasmic calcium induced by blue light. Plant Physiol. 133 1464–1470. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Frank, A. (1868). Beiträge zur Pflanzenphysiologie. (Leipzig, Germany: Engelmann).
  • Friml, J., Wisniewska, J., Benkova, E., Mendgen, K., and Palme, K. (2002). Lateral relocation of auxin efflux regulator PIN3 mediates tropism in Arabidopsis. Nature 415 806–809. [ PubMed ] [ Google Scholar ]
  • Gallanger, S., Short, T.S., Ray, P.M., Pratt, L.H., and Briggs, W.R. (1988). Light-mediated changes in two proteins found associated with plasma membrane fraction from pea stem sections. Proc. Natl. Acad. Sci. USA 85 8003–8007. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Galston, A.W. (1950). Riboflavin, light and the growth of plants. Science 111 619–624. [ PubMed ] [ Google Scholar ]
  • Galston, A.W., and Baker, R.S. (1949). Studies on the physiology of light action. II. The photodynamic action of riboflavin. Am. J. Bot. 36 773–780. [ PubMed ] [ Google Scholar ]
  • Gardner, G., Shaw, S., and Wilkins, M.B. (1974). IAA transport during the phototropic responses of intact Zea and Avena coleoptiles. Planta 28 439–478. [ PubMed ] [ Google Scholar ]
  • Gehring, C.A., Williams, D.A., Cody, S.H., and Parish, R.W. (1990). Phototropism and geotropism in maize coleoptiles are spatially correlated with increases in cytosolic free calcium. Nature 345 528–530. [ PubMed ] [ Google Scholar ]
  • Guillemin. (1858). Production de la chlorophylle et direction des tiges sous l'influence des rayons ultra-violets, calorifques et lumieux du spectra solaire. Ann. Sci. Nat. 4 154–172. [ Google Scholar ]
  • Haig, C. (1935). The phototropic responses of Avena in relation to intensity and wavelength. Biol. Bull. 69 305–324. [ Google Scholar ]
  • Hangarter, R.P. (1997). Gravity, light and plant form. Plant Cell Environ. 20 796–800. [ PubMed ] [ Google Scholar ]
  • Harper, R.M., Stowe-Evans, E.L., Luesse, D.R., Muto, H., Tatematsu, K., Watahiki, M.K., Yamamoto, K., and Liscum, E. (2000). The NPH4 locus encodes the auxin response factor ARF7, a conditional regulator of differential growth in aerial Arabidopsis tissue. Plant Cell 12 757–770. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Heslop-Harrison, J.S. (1980). Darwin and the movements of plants: A retrospect. In Plant Growth Substances, F. Skoog, ed (Berlin: Springer-Verlag), pp. 1–14.
  • Huala, E., Oeller, P.W., Liscum, E., Han, I.S., Larsen, E., and Briggs, W.R. (1997). Arabidopsis NPH1: A protein kinase with a putative redox-sensing domain. Science 278 2120–2123. [ PubMed ] [ Google Scholar ]
  • Iino, M., and Briggs, W.R. (1984). Growth distribution during first positive phototropic curvature of maize coleoptiles. Plant Cell Environ. 160 97–104. [ Google Scholar ]
  • Iino, M., Briggs, W.R., and Schafer, E. (1984). Phytochrome-mediated phototropism in maize seedling shoots. Planta 160 41–51. [ PubMed ] [ Google Scholar ]
  • Inada, S., Ohgishi, M., Mayama, T., Okada, K., and Sakai, T. (2004). RPT2 is a signal transducer involved in phototropic response and stomatal opening by association with phototropin 1 in Arabidopsis thaliana. Plant Cell 16 887–896. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Janoudi, A.K., Gordon, W.R., Wagner, D., Quail, P., and Poff, K.L. (1997. a). Multiple phytochromes are involved in red-light-induced enhancement of first-positive phototropism in Arabidopsis thaliana. Plant Physiol. 113 975–979. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Janoudi, A.K., Konjevic, R., Apel, P., and Poff, K.L. (1992). Time threshold for 2nd positive phototropism is decreased by a preirradiaiton with red-light. Plant Physiol. 99 1422–1425. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Janoudi, A.K., Konjevic, R., Whitelam, G., Gordon, W., and Poff, K.L. (1997. b). Both phytochrome A and phytochrome B are required for the normal expression of phototropism in Arabidopsis thaliana seedlings. Physiol. Plant. 101 278–282. [ Google Scholar ]
  • Janoudi, A.K., and Poff, K.L. (1990). A common fluence threshold for 1 st positive and 2 nd positive phototropism in Arabidopsis thaliana. Plant Physiol. 94 1605–1608. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Janoudi, A.K., and Poff, K.L. (1991). Characterization of adaptation in phototropism of Arabidopsis thaliana. Plant Physiol. 95 517–521. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Janoudi, A.K., and Poff, K.L. (1992). Action spectrum for enhancement of phototropism by Arabidopsis thaliana seedlings. Photochem. Photobiol. 56 655–659. [ Google Scholar ]
  • Janoudi, A.K., and Poff, K.L. (1993). Desensitization and recovery of phototropic responsiveness in Arabidopsis thaliana. Plant Physiol. 101 1175–1180. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Johnston, E.S. (1934). Phototropic sensitivity in relation to wavelength. Smithson. Misc. Coll. 92 1–17. [ Google Scholar ]
  • Kang, B.G., and Burg, S.P. (1974). Red Light enhancement of the phototropic response in pea stems. Plant Physiol. 53 445–448. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Khurana, J.P., and Poff, K.L. (1989). Mutants of Arabidopsis thaliana with altered phototropism. Planta 178 400–406. [ PubMed ] [ Google Scholar ]
  • Kogl, F., and Haagen-Smit (1931). I. Mitteilung uber pflanzliche wachstumsstoffe. Uber die vhemie des euchsstoffs. Proc. K. Ned. Akad. Wet. 34 1411–1416. [ Google Scholar ]
  • Kogl, F., and Schuringa, G.J. (1944). Uber die inaktivierung von auxin-a-lacton bei verschiedenen welenlagen und den einfluss von carotinoiden auf die licktreaktion. Hoppe-Seyler's Z. Physiol. Chem. 280 148–161. [ Google Scholar ]
  • Konjevic, R., Steinitz, B., and Poff, K.L. (1989). Dependence of the phototropic response of Arabidopsis on fluence rate and wavelength. Proc. Natl. Acad. Sci. USA 86 9876–9880. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Knieb, E., Salomon, M., and Rudiger, W. (2004). Tissue-specific and subcellular localization of phototropin deterimined by immuno-blotting. Planta 218 843–851. [ PubMed ] [ Google Scholar ]
  • Lasceve, G., Leymarie, J., Olney, M.A., Liscum, E., Christie, J.M., Vavasseur, A., and Briggs, W.R. (1999). Arabidopsis contains at least four independent blue-light-activated signal transduction pathways. Plant Physiol. 120 605–614. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Leong, T., and Briggs, W.R. (1982). Evidence from studies with acriflourfen for participation of a flavin-cytochrome complex in blue light phototropism for phototropism of oat coleoptiles. Plant Physiol. 70 875–881. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Liscum, E., and Briggs, W.R. (1995). Mutations in the NPH1 locus of Arabidopsis disrupt the perception of phototropic stimuli. Plant Cell 7 473–485. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Liscum, E., and Briggs, W.R. (1996). Mutations of Arabidopsis in potential transduction and response components of the phototropic signaling pathway. Plant Physiol. 112 291–296. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Liscum, E., and Reed, J.W. (2002). Genetics of Aux/IAA and ARF action in plant growth and development. Plant Mol. Biol. 49 387–400. [ PubMed ] [ Google Scholar ]
  • Liscum, E., Young, J.C., Poff, K.L., and Hangarter, R.P. (1992). Genetic separation of phototropism and blue light inhibition of stem elongation. Plant Physiol. 100 267–271. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Macleod, K., Digby, J., and Firn, R.D. (1985). Evidence inconsistent with the Blaauw model of phototropism. J. Exp. Bot. 36 312–319. [ Google Scholar ]
  • McKeon, R., ed. (1947). Introduction to Aristotle. (New York: The Modern Library).
  • Motchoulski, A., and Liscum, E. (1999). Arabidopsis NPH3: A NPH1 photoreceptor-interacting protein essential for phototropism. Science 286 961–964. [ PubMed ] [ Google Scholar ]
  • Munoz, V., and Bulter, W.L. (1975). Photoreceptor pigment for blue light in Neurospora crassa. Plant Physiol. 55 421–426. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Naqvi, S.M. (1972). Possible role of abcissic acid in phototropism. Z. Pflanzenphysiol. 67 454–456. [ Google Scholar ]
  • Noh, B., Bandyopadhyay, A., Peer, W.A., Spalding, E.P., and Murphy, A.S. (2003). Enhanced gravi- and phototropism in plant mdr mutants mislocalizing the auxin efflux protein PIN1. Nature 423 999–1002. [ PubMed ] [ Google Scholar ]
  • Overbeek, J.V. (1932). An analysis of phototropism in dicotyledons. Proc. K. Ned. Akad. Wet. 35 1325–1335. [ Google Scholar ]
  • Overbeek, J.V. (1933). Wuchsstoff, lichtwachstumsreaktion und phototropismus bei Raphanus. Rec. Trav. Bot. Neerland. 30 537–626. [ Google Scholar ]
  • Ovid, Melville, A.D., and Kenney, E.J. (1998). Metamorphoses. (Oxford, UK: Oxford University Press).
  • Paal, A. (1918). Uber phototropische reizleitung. Jahr. Wiss. Bot. 58 406–458. [ Google Scholar ]
  • Parker, K., Baskin, T.I., and Briggs, W.R. (1989). Evidence for a phytochrome-mediated phototropism in etiolated pea seedlings. Plant Physiol. 89 493–497. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Parks, B.M., Quail, P.H., and Hangarter, R.P. (1996). Phytochrome A regulates red-light induction of phototropic enhancement in Arabidopsis. Plant Physiol. 110 155–162. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Payer, J. (1842). Memoire sur la tendance des tiges ver la lumiere. C. R. Seances Acad. Sci. 813–816.
  • Pickard, B.G., and Thimann, K.V. (1964). Transport and distribution of auxin during tropistic response. II. The lateral migration of auxin in phototropism of coleoptiles. Plant Physiol. 39 341–350. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Plato. (2000). Timaeus. (Indianapolis, IN: Hackett Publishing Company).
  • Poggioli, S. (1817). Della influenza che ha il raggio magnetico sulla vegetatione delle piante. Opuscoli Scientifici (Bologna) I, 9–23.
  • Pringsheim, E.G. (1909). Studien zur heliotropischen stimmung und prasentationzeit. Beitr. Biol. Pflanzen. 9 415–478. [ Google Scholar ]
  • Pringsheim, E.G. (1912). Die Reizbewegungen der Planzen. (Berlin: Springer-Verlag).
  • Ray, J., Camel, G.J., and Tournefort, J.P. (1686). Historia plantarum. (London: Henricum Faithorne).
  • Reymond, P., Short, T.W., Briggs, W.R., and Poff, K.L. (1992. a). Blue light activates a specific protein kinase in higher plants. Plant Physiol. 100 655–661. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Reymond, P., Short, T.W., Briggs, W.R., and Poff, K.L. (1992. b). Light-induced phosphorylation of a membrane protein plays an early role in signal transduction for phototropism in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 89 4718–4721. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Rich, T.C.G., Whitelam, G., and Smith, H. (1985). Phototropism and axis extension in light-grown mustard ( Sinapis alba L.) seedlings. Photochem. Photobiol. 42 789–792. [ Google Scholar ]
  • Rothert, W. (1894). Uber heliotropismus. Beitr. Biol. Pflanzen. 7 1–211. [ Google Scholar ]
  • Sachs, J. (1864). Wirkungen farbigen lichts auf pflanzen. Bot. Zeitung 47 353–358. [ Google Scholar ]
  • Sachs, J., Garnsey, H.E.F., and Balfour, I.B. (1890). History of Botany (1530–1860). (Oxford, UK: Clarendon Press).
  • Sakai, T., Kagawa, T., Kasahara, M., Swartz, T.E., Christie, J.M., Briggs, W.R., Wada, M., and Okada, K. (2001). Arabidopsis nph1 and npl1: Blue light receptors that mediate both phototropism and chloroplast relocation. Proc. Natl. Acad. Sci. USA 98 6969–6974. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Sakai, T., Wada, T., Ishiguro, S., and Okada, K. (2000). RPT2. A signal transducer of the phototropic response in Arabidopsis. Plant Cell 12 225–236. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Sakamoto, K., and Briggs, W.R. (2002). Cellular and subcellular localization of phototropin 1. Plant Cell 17 1723–1735. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Salomon, M., Christie, J.M., Knieb, E., Lempert, U., and Briggs, W.R. (2000). Photochemical and mutational analysis of the FMN-binding domains of the plant blue light receptor, photoropin. Biochemistry 39 9401–9410. [ PubMed ] [ Google Scholar ]
  • Salomon, M., Zacherl, M., and Rudiger, W. (1997). Asymmetric, blue light-dependent phosphorylation of a 116-kilodalton plasma membrane protein can be correlated with the first- and second-positive phototropic curvature of oat coleoptiles. Plant Physiol. 115 485–491. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Schmidt, W., Hart, J., Filner, P., and Poff, K.L. (1977). Specific inhibition of phototropism in corn seedlings. Plant Physiol. 60 736–738. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Shank, A. (1950). Phototropism? Annu. Rev. Plant Physiol. 1 59–74. [ Google Scholar ]
  • Sharrock, R. (1672). The History of the Propagation and Improvement of Vegetables. (Oxford, UK: W. Hall).
  • Shemp, J. (1947). Plants in Plato's Timaeus. Class. Q. 41 53–60. [ Google Scholar ]
  • Shen-Miller, J., Cooper, P., and Gordon, S.A. (1969). Phototropism and photoinhibiiton of basipolar transport of auxin in oat coleoptiles. Plant Physiol. 44 491–496. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Shen-Miller, J., and Gordon, S.A. (1966). Hormonal relations in the phototropic response. IV. Light-induced changes of endogenous auxins in the coleoptile. Plant Physiol. 41 831–841. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Short, T.W., and Briggs, W.R. (1990). Characterization of a rapid, blue light-mediated change in detectable phosphorylation of a plasma meembrane protein from etiolated pea ( Pisum sativum L.) seedlings. Plant Physiol. 92 179–185. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Short, T.W., Porst, M., and Briggs, W.R. (1992). A photoreceptor system regulating in vivo and in vitro phosphorylation of a pea plasma-membrane protein. Photochem. Photobiol. 55 773–781. [ Google Scholar ]
  • Silk, W.K. (1984). Quantitative descriptions of development. Annu. Rev. Plant Physiol. 35 479–518. [ Google Scholar ]
  • Steinitz, B., and Poff, K.L. (1986). A single positive phototropic response induced with pulsed-light in hypocotyls of Arabidopsis thaliana seedlings. Planta 168 305–315. [ PubMed ] [ Google Scholar ]
  • Stowe-Evans, E.L., Luesse, D.R., and Liscum, E. (2001). The enhancement of phototropic curvature via photoreversible phytochrome A-dependent modulation of auxin responsiveness. Plant Physiol. 126 826–834. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Tatematsu, K., Kumagai, S., Muto, H., Sato, A., Watahiki, M.K., Harper, R.M., Liscum, E., and Yamamoto, K.T. (2004). MASSUGU2 encodes Aux/IAA19, an auxin-regulated protein that functions together with the transcriptional activator NPH4/ARF7 to regulate differential growth responses of hypocotyl and formation of lateral roots in Arabidopsis thaliana. Plant Cell 16 379–393. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Theophrastus. (1976). De Causis Plantarum. (London: Harvard University Press).
  • Vierstra, R.D., and Poff, K.L. (1981. a). Mechanism of specific inhibition of phototropism by phenylacetic acid in corn seedling. Plant Physiol. 67 1011–1015. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Vierstra, R.D., and Poff, K.L. (1981. b). Role of carotenoids in the phototropic response of corn seedlings. Plant Physiol. 68 798–801. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • von Erhardt-Siebold, E. (1937). The heliotrope tradition. Osiris 3 22–46. [ Google Scholar ]
  • von Wiesner, J. (1878). Die Heliotropischen Erscheinungen im Pflanzenreiche: Eine Physiologische Monographie. (Wien, Germany: Kaiserlich-königlichen Hof- und Staatsdruckerei).
  • Wald, G., and Du Buy, H.G. (1936). Pigments of the oat coleoptile. Science 84 247. [ PubMed ] [ Google Scholar ]
  • Webster, C. (1966). The recognition of plant sensitivity by English botanists in the seventeenth century. Isis 57 5–23. [ Google Scholar ]
  • Went, F.W. (1926). On growth accelerating substances in the coleoptile of Avena sativa. Proc. K. Akad. Wet. 30 10–19. [ Google Scholar ]
  • Went, W.F. (1928). Wuchsstoff und Wachstum. Recl. Trav. Bot. Neerland. 25 1–116. [ Google Scholar ]
  • Went, F.W., and Thimann, K.V. (1937). Phytohormones. (New York: Macmillan Company).
  • Whippo, C.W., and Hangarter, R.P. (2003). Second positive phototropism results from coordinated co-action of the phototropins and cryptochromes. Plant Physiol. 132 1499–1507. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Whippo, C.W., and Hangarter, R.P. (2004). Phytochrome modulation of blue-light phototropism. Plant Cell Environ. 27 1223–1228. [ Google Scholar ]
  • Whippo, C.W., and Hangarter, R.P. (2005). A brassinosteroid-hypersensitive mutant of BAK1 indicates that a convergence of photomorphogenic and hormonal signaling modulates phototropism. Plant Physiol. 139 448–457. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Zantedeschi, M. (1843). De l'influence qu'exercent sur la vegetation de plants et la germination des graines les rayons transmis a travers des verres colores. C. R. Seances Acad. Sci. 747–749.

Student Sheet 8 – Phototropism: the Response of Seedlings to Light

  • Share to Facebook
  • Share to Twitter
  • Share via Email

Age Ranges:

experiment for phototropism

Seedlings growing on a windowsill will often bend towards the window as they respond to light – phototropism. But what exactly are the seedlings responding to? Which wavelengths of light stimulate the phototropic response?

The technique for this experiment helps students design an investigation to find out more about this tropic response. Students germinate seedlings in enclosed containers, with a coloured filter over the small hole allowing light in. Students predict which seedlings will demonstrate phototropism, and which will remain unaffected.

The students’ sheet contains a number of suggestions for further experiments, while the  worksheet outlines the basic technique.

Download the student sheets and teachers’ notes from the links on the right.

What's included?

  • SAPS Sheet 8 - The response of seedlings to light - Student Notes
  • SAPS Sheet 8 - The response of seedlings to light - Student Sheet
  • SAPS Sheet 8 - The response of seedlings to light - Technical and Teaching Notes
  • Plant growth
  • Plant reproduction
  • Plant responses

Related content

Teaching resources.

  • Investigating Thigmotropism
  • Tackling tropisms: gravitropism and phototropism

Exercises and Materials for Teaching Math

Plant Phototropism Experiment For Students

Phototropism experiment for students

  • Spray bottle
  • 8 bean seeds
  • 8 small pots
  • Two 1-foot-tall cardboard boxes with lids
  • Masking tape
  • Box cutter knife (with an adult to help)
  • Piece of cardboard
  • 2 small lamps
  • 2 full spectrum light bulbs
  • One 3” x 3” piece of clear, red green and blue cellophane
  • Plant two of the bean seeds in two of the pots, give them water and let them grow. Wait until you can see them start to poke out of the soil.
  • Use the waiting time to prepare your boxes. With each box, cut a hole 2” in diameter about 3 in. from the bottom. Take the clear cellophane and put it over the hole on one box. On the other box, place red cellophane over the hole. Now, one box will let every kind of light in and one box will only let red light in.
  • Place your plants in the boxes, one in each. Use the ruler to place them 2 in. away from the hole. Use your camera to take a picture of the plants, looking down from above.
  • Put the boxes on different sides of one room.
  • Place the lamps next to the boxes, one for each, on the side with the cellophane covered hole. Make sure the lamps are equal distance from the holes, so everything is the same.
  • Cover the boxes.
  • In the morning, turn on the lamp. In the night, turn off the lamp. Repeat for a week.
  • After the week, uncover the boxes and take a picture from above again. Then take the plants out and take a picture from the front. Did the plants grow differently?
  • Repeat the experiment, but this time, change the red cellophane to blue cellophane. Then, repeat with green cellophane.
  • Look at the pictures you’ve taken of each plants. Did the plants turn towards one color more than another? Did they not turn towards one color at all?

Results The plants in the box with the white cellophane, or the control plants, will have done better. The plants growing in red and blue light will have done better than the one in green light, the plants will grow towards both red and blue but will not have grown towards the green light.

1st Grade Science

1st grade science

Games, Quizzes, Worksheets

Go here >>>

2nd Grade Science

2nd grade science

3rd Grade Science

3rd grade science

Games, Quizzes, Worksheets.

4th Grade Science

4th grade science

We offer PDF sheets in the highest quality.

  • 1st grade science
  • 2nd grade science
  • 3rd grade science & more

E-learning science games

  • Pirate Games
  • Moonshot Game
  • Catapult Game
  • and More More Games.

We make our science lessons appealing to kids by blending activities with interesting science puzzles. Track and find answers .

We cover topics on: weather, animals, ecosystems, life cycles, laboratory readings, fossils, plants, adaptation, force, mixtures, classification etc.

We help parents and teachers supplement their kid's science course. Our resources are therefore free and can be shared online.

Science practice for kids

  • Science Worksheets
  • Science Games
  • Science Puzzles
  • Science Quizzes
  • Science Articles

PRINTABLE EXERCISES

  • Identify ecosystems
  • Compare ecosystems
  • Digestive system
  • Earth processes & More

Science Branches

  • Physical Science
  • Environmental science
  • Life science

MyQuestionIcon

Demonstrate experiment to show A phototropism B.geotropism C.chemotropism D.hydrotropism

1. demonstration of phototropism in stem: take a potted seedling of a plant. place it in heliotropic (phototropic) chamber with a small hole on one side (). close the top of the chamber. place it in a position that allows sufficient light to pass through the hole. keep it there for a couple of days. open the chamber and observe the seedling. it shows a bending (curvature) of the stem towards the hole (direction of light stimulus). this response is positive phototropism. 2. demonstration of geotropism in a plant: take a potted seedling. place it horizontally within a dark wooden chamber (). open the chamber after one day. the stem will be seen bent upwards, i.e., in a direction opposite to the pull of gravity. this is negative geotropism. take the roots out of the pot. they will show a curvature or bend opposite to the bend of the stem. this curvature is in the direction of the pull of gravity. it is positive geotropism. chemotropism: certain chemical substances are responsible to bring about curvature movements in plant organs. for instance, movement of pollen tube towards ovary due to absorption of calcium and borate from style of carpel; movement of tentacles in drosera, closing of lid of nepenthes due to nitrogenous food, and penetration of haustoria of parasite into host body etc. hydrotropism: the paratonic curvature movements of growth in relation to the stimulus of water are called hydrotropic movements. the tropic response to the stimulus of water is called hydrotropism. the roots show positive hydrotropic response, i.e., they bend towards the water ). hydrotropism is stronger in roots compared to geotropism..

flag

Question 5 Design an experiment to demonstrate hydrotropism.

thumbnail

From “Human-Centered” to “System-Oriented”: Eco-Cultural Legacy of Feng Shui and Scientific Principles for Establishing Modern Resilient Cities

  • First Online: 28 August 2024

Cite this chapter

experiment for phototropism

  • Wenjian Pan 6  

Not only are cities confronting risks from global warming, but there are also multiple ecological challenges to achieving sustainable development goals (SDGs), particularly increasing urban heat island effects, more frequent heat waves, and the threat of flood hazards. However, blue-green infrastructure (BGI) plays a crucial role in mitigating these risks while maintaining the normal operation of cities. Feng Shui, a theorization of Chinese ancients’ habitation experiences, offers an approach to interacting with nature and complying with the laws of natural eco-systems. It has provided many philosophical thoughts and scientific guidance on making habitats resilient in the face of environmental changes and multi-hazards. Using Hong Village as a standard application model for Feng Shui, this chapter reinterprets the naturally adaptable principles in Feng Shui and analyzes how its passive strategies can be adopted in dwelling systems. It finds that Feng Shui advocates co-developing with nature and treats human habitats as an organic living system that is interlinked with surrounding environments. Specifically, it emphasizes flowing air ( Feng ) and flowing water ( Shui ) as two fundamental natural elements to be integrated into construction activities, which can contribute to the accumulation and circulation of energies, resources, and dynamics. As such, a balance between humans’ activities and the natural processes is critical in Feng Shui, making it a useful philosophy for this context. Based on these reinterpretations, the present chapter establishes a multi-scalar framework for scientific modeling and modern applications. Ultimately, the Cheonggyecheon urban restoration project in South Korea and the Yanweizhou ecological rehabilitation project in China are taken as two successful modern cases to showcase integrations of Feng Shui principles into BGIs in contemporary cities, and they aim to adapt to the emerging challenges of increasing urban heat and flood hazards, respectively. To conclude, a few nature-based, and soft intervention-oriented approaches are outlined for the construction and rehabilitation of BGIs to strengthen urban resilience and achieve the long-term goals of sustainable development.

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

Access this chapter

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Durable hardcover edition
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Abbreviations

Wind/Flowing air

The Book of Change

Geomancy/Deep investigation of land

Compass School

Chinese compass

Man can conquer nature

Water inlet/Water gap

Rainwater flows from the pitched roofs to the courtyard

Tai Chi/The Great ultimate

Patio/Light well

Integration of humans and nature

Five elements

Form School

Lair, cave, and hole/construction site

Light/Positive dimension

Connecting to mountainous terrain and water areas

Darkness/Negative dimension

Adopting local methods and approaches

Facing south and having mountains at the back

Al, S. (2022). Multi-functional urban design approaches to mange floods: Examples from Dutch cities. Journal of Urban Design, 27 (2), 270–278.

Article   Google Scholar  

Alberti, M. (2009). Advances in Urban Ecology: Integrating Humans and Ecological Processes in Urban Ecosystems . Springer.

Google Scholar  

Almodovar-Melendo, J., & Cabeza-Lainez, J. (2018). Environmental features of Chinese architectural heritage: The standardization of form in the pursuit of equilibrium with nature. Sustainability, 10 , 2443.

Article   CAS   Google Scholar  

Architecture 2030. (2022). Why the built environment? Retrieved August 15, 2023, from https://architecture2030.org/why-the-building-sector/ .

Bettencourt, L. M. A. (2021). Introduction to Urban Science: Evidence and Theory of Cities as Complex Systems . MIT Press.

Bhadouria, R., Tripathi, S., Singh, P., Joshi, P. K., & Singh, R. (2023). Urban metabolism and global climate change: An overview.  Urban Metabolism and Climate Change: Perspective for Sustainable Cities , 3–22.

Browning, W., Ryan, C., & Clancy, J. (2014). 14 Patterns of Biophilic Design: Improving Health & Well-being in the Built Environment . Terrapin Bright Green LLC.

Bruun, O. (2008). An introduction to Feng Shui . Cambridge University Press.

Book   Google Scholar  

Burki, T. (2023). Health at COP28: Three things to watch for. The Lancet, 402 (10416), 1954.

Chen, X. Y., Zheng, B., & Fu, X. Z. (2013). Actual measurement analysis of lowering temperature in cooling alley of folk house. Architectural Journal, 02 , 82–85.

Chinese Traditional Village Museum. (2020). Hong Cun (Hong village) . (Accessed: http://main.dmctv.com.cn/villages/34102310101/Index.html ).

Chung, J. H., Hwang, K. Y., & Bae, Y. K. (2012). The loss of road capacity and self-compliance: Lessons from the Cheonggyecheon stream restoration. Transport Policy, 21 , 165–178.

Chung, M. G., Frank, K. A., Pokhrel, Y., Dietz, T., & Liu, J. G. (2021). Natural infrastructure in sustaining global urban freshwater ecosystem services. Nature Sustainability, 4 , 1068–1075.

Cohen-Shacham, E., Walters, G., Janzen, C., & Maginnis, S. (Eds.). (2016). Nature-based Solutions to address global societal challenges (xiii + 97pp.). Gland, Switzerland: IUCN.

Coseo, P., & Larsen, L. (2015). Cooling the heat island in compact urban environments: The effectiveness of Chicago’s green alley program. Procedia Engineering, 118 , 691–710.

Deng, Y., Yoshihiko, Y., & Kim, G. (2006). Urban space transformation informed by the naming changes of the Korean capital city. Beijing Planning Review , (04), 42–46.

Ding, Y. (2011). Kan Yu: A Forgotten Intangible Heritage in China . Southeast University Press.

Duan, J., & Jie, M. H. (2009). Spatial Analysis of the Ancient Village of Hongcun World Cultural Heritage . Southeast University Press.

Gao, Y. F., Cheng, J. J., & Wang, Z. W. (2007). Ecology physical environment computational simulation of ideal Feng Shui position model. Building Science, 23 (6), 19–23.

Geddes, P. (1923). The Valley Section from Hills to Sea . New York City (United States).

Groffman, P. M., et al. (2017). Moving towards a new urban systems science. Ecosystems, 20 , 38–43.

Guattari, F. (2000). The three ecologies . London; New Brunswick, N.J.: Athlone Press.

Guo, X. N., Cheng, J., Yin, C. L., Li, Q., Chen, R. S., & Fang, J. Y. (2023a). The extraordinary Zhengzhou flood of 7/20, 2021: How extreme weather and human response compounding to the disaster. Cities, 134 , 104168.

Guo, Z. Y., Zheng, J. X., Li, S. Z., & Zhou, X. N. (2023b). Orientation of one health development: Think globally and act locally. Science in One Health, 2 , 100042.

Guo, H. Y. (2023). Analysis and reflection on the difficulties of modern inheritance of chinese traditional wooden architecture. New Architecture , December 7.

Hauck, T., Keller, R., & Kleinekort, V. (2012). Infrastructure Urbanism: Addressing the in-between .

Jie, M. H. (2006). The ancient village in Huizhou: The research on influential factors of Hongcun village spatial structure. Journal of Shanghai City Management Vocational and Technical College , A42–A45.

Jin, Y. H. (2005). Hong Cun (Hong village) . Jiangsu Education Publishing House.

Kang, L. (1999). Feng Shui and architecture . Baihua Literature and Art Publishing House.

Kapoor, V., Tripathi, S., Devi, R. S., Srivastava, P., & Bhadouria, R. (2020). Chapter 6–ecological economics of an urban settlement: an overview. Urban Ecology , 91–110.

La Loggia, G., Puleo, V., & Freni, G. (2020). Floodability: A new paradigm for designing urban drainage and achieving sustainable urban growth. Water Resources Management, 34 , 3411–3424.

Li, Y. Q., & Wang, C. L. (2019). Yanweizhou ecological embankment strategy and its effects on flood control in the Jinhua River Basin. Acta Ecologica Sinica, 39 (16), 5955–5966.

Liu, Y. S. (2000). Shui Kou: Huizhou Minju de Zezhi Guannian (Water gap: a concept of site selection for residential buildings). Wenshi Zhishi , 76–80.

Luo, Y., & He, X. X. (2008). Zhongguo Feng Shui Shi (The history of Chinese Feng Shui) . Jiuzhou Publishing House.

Mak, M. Y., & Ng, S. T. (2005). The art and science of Feng Shui–a study on architects’ perception. Building and Environment, 40 (3), 427–434.

Mak, M. Y., & So, A. T. (2011). Scientific Feng Shui for the Built Environment: Fundamentals and Case Studies . City University of Hong Kong Press.

Marco, E., & Burgess, S. (2015). Chapter 28-healthy housing. In The routledge Handbook of Planning for Health and Well-being: Shaping a Sustainable and Healthy Future (pp. 403–416). London; New York: Routledge, Taylor & Francis Group.

Margolis, L., & Robinson, A. (2007). Living Systems: Innovative Materials and Technologies for Landscape Architecture . Birkhäuser Architecture.

McHarg, I. L. (1995). Design with nature . Wiley.

McPhearson, T., Raymond, C. M., et al. (2021). Radical changes are needed for transformations to a good Anthropocene. Npj Urban Sustainability , 5 , 1–13.

Mostafavi, M. (Ed.). (2010). Ecological urbanism . Lars Müller Publishers.

Ng, E. (Ed.). (2010). Designing High-density Cities for Social and Environmental Sustainability . London Sterling.

Ou, Y. W. (2001). Simple analysis on features of courtyard space structure in Huizhou traditional residences. Journal of Beijing Institute of Civil Engineering and Architecture, 15–17.

Pan, W. J. (2019). Diverse Environmental Performances of Urban Villages and Insights for Enhancing Quality of Urban Renewal in Shenzhen . The University of Hong Kong.

Pan, W. J. (2023). Self-adpative hybrid urban morphologies community (HUMC): Its shared environment and soft intervention for sustainable urban governance. Building and Environment, 236 , 110251.

Pan, W. J., & Du, J. (2021). Towards sustainable urban transition: A critical review of strategies and policies of urban village renewal in Shenzhen, China. Land Use Policy, 111 , 105744.

Pan, M. W. (2004). Buffalo-shaped village in Southern Anhui Yi County—Totem and Feng Shui in the planning of ancient Hong village. Journal of Kunming University of Science and Technology (Science and Technology), 29 (2), 103–106.

Park, C. Y., Lee, D. K., Asawa, T., Murakami, A., Kim, H. G., Lee, M. K., & Lee, H. S. (2019). Influence of urban form on the cooling effect of a small urban river. Landscape and Urban Planning, 183 , 26–35.

Paton, M. J. (2021). Science and Fengshui : The concept Shi, rationality and emotion, and the ritualisation of knowledge. Science & Education, 30 , 1371–1386.

Pickett, S. T. A., Cadenasso, M. L., Childers, D. L., McDonnell, M. J., & Zhou, W. Q. (2016). Evolution and future of urban ecological science: Ecology in , of , and for the city. Ecosystem Health and Sustainability, 2 (7), e01229.

Pickett, S. T. A., Frantzeskaki, N., Andersson, E., Barau, A. S., Childers, D. L., Hoover, F., Lugo, A. E., McPhearson, T., Nagendra, H., Schepers, S., & Sharifi, A. (2024). Shifting forward: Urban ecology in perspective. Ambio, 53 , 890–897.

Porkert, M. (1978). Theoretical Foundations of Chinese Medicine–Systems of Correspondence . The MIT Press.

Robinson, T., & Ji, M. (2022). Freeway removal: From Cheonggyecheon to Seoullo 7017. In Sustainable, smart and solidary Seoul: Transforming an asian megacity (pp. 39–71). Springer.

Romanello, M., et al. (2023). The 2023 report of the Lancet Countdown on health and climate change: The imperative for a health-centred response in a world facing irreversible harms. The Lancet, 402 (10419), 2346–2394.

Ru, J. F. (1993). Theoretical analysis of “Cave” in Feng Shui from the perspective of architectural ecology. Time+Architecture, 1 , 20–22.

Salmond, J. A., Roth, M., Oke, T. R., Christen, A., & Voogt, J. A. (2012). Can surface-cover tiles be summed to give neighborhood fluxes in cities? Journal of Applied Meteorology and Climatology, 51 (1), 133–149.

Seoul Development Institure. (2003). Feasilibility Study and Master Plan of Cheonggyecheon (Non-technical part summary report) final report . Cheonggyecheon Restoration Research Center, Seoul, Korea (in Korean).

Shan, D. Q. (2009). Anhui Min Ju (Vernacular architecture of Anhui Province) . China Architecture & Building Press.

Shou, T., & Xue, D. (2022). Resilient containers: An exploration into the tectonic logics of the water environment regulation in Huizhou’s traditional dwellings. Architectural Journal, 12 , 84–91.

Shou, T., Zhang, T., & Liu, Q. (2019). The research of architectural thermodynamic model of Huizhou vernacular architecture. Jian Zhu Shi (Architects), 6 , 35–44.

Sun-Kee, H. C., Song, I. J., & Wu, J. G. (2007). Fengshui theory in urban landscape planning. Urban Ecosystems, 10 (3), 221–237.

Tang, L., Nikolopoulou, M., Zhao, F. Y., & Zhang, N. (2012). CFD modeling of the built environment in Chinese historic settlements. Energy and Buildings, 55 , 601–606.

Turenscape. (2017). Resilient landscape–Yanweizhou Park, Jinhua. Urban Environment Design , (03), 327–329.

Turner, V. K., Middel, A., & Vanos, J. K. (2023). Shade is an essential solution for hotter cities. Nature, 619 , 694–697.

UNEP (United Nations Environment Programme). (2021). International Principles of Good Practice for Sustainable Infrastructure: Integrated, System-level Approaches for Policymakers . https://wedocs.unep.org/20.500.11822/34853 .

UNEP (United Nations Environment Programme). (2023). Nature-based Infrastructure: How Natural Infrastructure Solutions Can Address Sustainable Development Challenges and the Triple Planetary Crisis . https://wedocs.unep.org/20.500.11822/44022 .

United Nations. (1987). Report of the World Commission on Environment and Development: Our Common Future .

United Nations. (2022). The Sustainable Development Goals Report 2022 . Casemate Publishers and Book Distributors, LLC.

Wang, D. G., Lyu, Q. Y., Wu, Y. F., & Fan, Z. Q. (2019). The characteristic of regional differentiation and impact mechanism of architecture style of traditional residence. Journal of Natural Resources, 34 (9), 1864–1885.

Wang, Q. H. (2005). Feng Shui Lilun Yanjiu (Study of Feng Shui theory) . Tianjin University Press.

Wang, X. B., & Jia, D. (2011). The analysis of contemporary on constructing strategy in traditional settlements and architecture space: The thoughts of water system of landscape and architecture space changes in Huizhou Hongcun. Huazhong Architecture , 83–85.

Welter, V. M. (2002). Biopolis: Patrick Geddes and the city of life . MIT Press.

Wilhelm, H., & Wilhelm, R. (1995). Understanding the I Ching: The Wilhelm Lectures on the Book of Changes . Princeton University Press.

Wong, C. (2023). Earth just had its hottest year on record–climate change is to blame. Nature , 10 November.

Wu, L. Y. (2014). Zhongguo Minju Shi (A history of Chinese vernacular dwellings) . China Architecture & Building Press.

Wu, L. Y. (2017). Development tendency of human settlements sciences. Journal of Urban and Regional Planning, 9 (02), 1–14.

CAS   Google Scholar  

Xia, Y., Guo, J., & Zou, Y. (2022). Preservation strategy and soundscape survey of tourism-oriented traditional villages: Taking Hongcun as an example. Huazhong Architecture, 40 (10), 173–177.

Xu, J., & Chung, C. K. L. (2014). “Environment” as an evolving concept in China’s urban planning system. International Development Planning Review, 36 (4), 391–412.

Yan, M. K., Gu, Y., & Lyu, L. S. (2023). A review of bionics applications in urban planning. Huazhong Architecture, 41 (05), 11–16.

Yoon, H. K. (1989). The theory on the origin and development of ancient Chinese geomancy (Feng-Shui). Studies in the History of Natural Science, 8 (1), 84–89.

You, X. Y. (2023). Flood-prone areas are hotspots for urban development. Nature , October 04.

Yu, G., Gan, H. G., & Guo, Y. Q. (2023). Location and fortune: An exploration of the Buddhism and Daoism roles of geomancy in the song dynasty. Religions, 14 , 859.

Yu, K. J., Li, H. L., & Li, D. H. (2008). The negative approach and ecological infrastructure: The smart preservation of natural systems in the process of urbanization. Journal of Natural Resources, 23 (6), 937–958.

Zhang, J. W. (2004). A Translation of the Ancient Chinese: The Book of Burial (Zang Shu) by Guo Pu (276–324) . Edwin Mellen Press.

Zhang, Z. Y. (2019). The revival practice and spatial planning inspiration of Cheonggyecheon River in South Korea. Urban Planning Forum, 06 , 120.

Zhong, W. Z., Xiao, W., & Zhang, T. (2023). Numerical investigations on natural ventilation in atria of China’s Southern Yangtze vernacular dwellings. Sustainable Cities and Society, 89 , 104341.

Download references

Acknowledgements

This chapter was partially developed from the author’s individual project on urban ecologies and environmental sustainability during his doctoral research in the Department of Architecture and working in the Urban Ecologies Design Lab at the The University of Hong Kong. The author thanks for Prof. Juan Du and Prof. Cole Roskam for the discussions at the initial stage of the study. Parts of the contents in this chapter have been presented in the 18th Multi-Hazards Symposium 2023 at Nanyang Technological University, Singapore.

Author information

Authors and affiliations.

School of Architecture and Urban Planning, Huazhong University of Science and Technology, Luoyu Road 1037, Hongshan District, Wuhan, 430074, China

Wenjian Pan

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Wenjian Pan .

Editor information

Editors and affiliations.

School of Environmental Sciences, Jawaharlal Nehru University, New Delhi, Delhi, India

P. K. Joshi

Department of Botany, University of Delhi, New Delhi, Delhi, India

Delhi College of Arts and Commerce, University of Delhi, New Delhi, Delhi, India

Rahul Bhadouria

Upadhyaya College, Dept of Botany, University of Delhi, Deen Dayal, New Delhi, Delhi, India

Sachchidanand Tripathi

Department of Botany, Panjab University, Chandigarh, Punjab, India

Rishikesh Singh

Rights and permissions

Reprints and permissions

Copyright information

© 2024 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Pan, W. (2024). From “Human-Centered” to “System-Oriented”: Eco-Cultural Legacy of Feng Shui and Scientific Principles for Establishing Modern Resilient Cities. In: Joshi, P.K., Rao, K.S., Bhadouria, R., Tripathi, S., Singh, R. (eds) Blue-Green Infrastructure for Sustainable Urban Settlements. Springer, Cham. https://doi.org/10.1007/978-3-031-62293-9_7

Download citation

DOI : https://doi.org/10.1007/978-3-031-62293-9_7

Published : 28 August 2024

Publisher Name : Springer, Cham

Print ISBN : 978-3-031-62292-2

Online ISBN : 978-3-031-62293-9

eBook Packages : Earth and Environmental Science Earth and Environmental Science (R0)

Share this chapter

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

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research

IMAGES

  1. Phototropism: Charles Darwin's & Boysen Jensen's Experiment

    experiment for phototropism

  2. Phototropic phototropism hi-res stock photography and images

    experiment for phototropism

  3. Phototropism

    experiment for phototropism

  4. Phototropism experiment.

    experiment for phototropism

  5. PPT

    experiment for phototropism

  6. Positive Phototropism Experiment Timelapse

    experiment for phototropism

COMMENTS

  1. Phototropism Experiment

    If the experiment worked correctly, you should have noticed that the seedlings that were covered with caps at the tip grew straight up, while the control seedlings and the seedlings with the bases covered bent towards the light. This is phototropism in action. Darwin correctly concluded that plants are able to "see" light using the tips of ...

  2. How Plants Respond To Light With Movement

    This diagram depicts early experiments with phototropism. Darwin showed that the phototropic response was lost when the tip of the emerging plant was cut off or covered with an opaque cap. Covering the base of the stem did not block phototropic bending. Dutch plant physiologist Boysen-Jensen showed in 1913 that placing a block of gelatin ...

  3. Plant Phototropism Experiment

    Cut a hole 2" in diameter about 3 inches from the bottom of each box. Place the clear cellophane over the hole. This will let all of the light into the box. Over the hole in the other box, place the red cellophane. This will only let red light into the box. Put one plant in the first box and one in the second.

  4. Phototropism: Bending towards Enlightenment

    When Robert Sharrock (1630-1684) repeated Browne's experiment, he concluded that the response was stimulated by fresh air and caused by growth rather than a mechanical consequence of wilting (Sharrock, 1672). These early phototropism experiments are significant because they provided some of the first scientific examples of plant sensitivity.

  5. Phototropism

    The presence of light is the most important factor in phototropism. Light energy stimulates the rate of photosynthesis in plants causing food production and generation of energy in the form of high-energy phosphate molecule, ATP. Phototropism is thus a survival strategy adopted by plants to maximize their light-harvesting capacity.

  6. Investigating Phototropism & Geotropism

    Phototropism in plant shoots. Plant shoots are positively phototropic, meaning they grow towards light; This ensures they maximise the amount of light they can absorb for photosynthesis; In the late 19th century and early 20th century, various experiments were conducted to investigate the process by which phototropism occurs. Many of the experiments were conducted using coleoptiles (a sheath ...

  7. Khan Academy

    Khanmigo is now free for all US educators! Plan lessons, develop exit tickets, and so much more with our AI teaching assistant.

  8. Understanding phototropism: from Darwin to today

    Jennifer J. Holland, Diana Roberts, Emmanuel Liscum, Understanding phototropism: from Darwin to today, Journal of Experimental Botany, Volume 60, Issue 7, May 2009, Pages 1969-1978, ... (1911) was able to gain further insight into Darwin's 'influence' in an experiment that used pieces of mica to disrupt the proposed influence's flow, and ...

  9. Phototropism

    In biology, phototropism is the growth of an organism in response to a light stimulus. Phototropism is most often observed in plants, but can also occur in other organisms such as fungi. The cells on the plant that are farthest from the light contain a hormone called auxin that reacts when phototropism occurs.

  10. American Journal of Botany

    Abstract. Light is a key environmental factor that drives many aspects of plant growth and development. Phototropism, the reorientation of growth toward or away from light, represents one of these important adaptive processes. Modern studies of phototropism began with experiments conducted by Charles Darwin demonstrating that light perception ...

  11. Phototropism

    Phototropism is the ability of a plant, or other photosynthesizing organism, to grow directionally in response to a light source. Plants and other autotrophs need to manufacture their own food; they usually do this through photosynthesis. Through photosynthesis, organisms convert water, carbon dioxide (CO 2) and light into sugars, which are ...

  12. Phototropism Definition, Mechanism & Examples

    Phototropism results in the plant growing towards the light, usually resulting in the bending of the entire plant. Much of a plant's photosynthesis occurs in leaves, and it is essential for new ...

  13. Khan Academy

    Khan Academy

  14. Phototropism and Geotropism

    The experiment on phototropism and geotropism focuses on understanding how plants respond to external stimuli, specifically light and gravity, and how these responses impact their growth and orientation. Phototropism investigates how plants bend or grow towards a source of light, emphasising their ability to optimise photosynthesis for energy ...

  15. Phototropism: Growing towards an Understanding of Plant Movement

    Abstract. Phototropism, or the differential cell elongation exhibited by a plant organ in response to directional blue light, provides the plant with a means to optimize photosynthetic light capture in the aerial portion and water and nutrient acquisition in the roots. Tremendous advances have been made in our understanding of the molecular ...

  16. Experiments to Show Phototropism (A-level Biology)

    Phototropism Experiment. We can investigate phototropism in plants using the following method. This will allow us to see the response of plants to light. Use 9 plant shoots. Plant all the shoots in individual plant pots, with the same soil type in each pot. Ensure that all the shoots are roughly the same height. Wrap some of the shoots in foil.

  17. Phototropism

    The following experiments were performed by different scientists in order to explain the mechanism of phototropism. Charles Darwin's Experiment In 1880, Charles Darwin, with the help of his son researched phototropism on canary grass and oat coleoptiles and recorded his observations in the book 'The Power of Movement in Plants'.

  18. Phototropism: Bending towards Enlightenment

    These early phototropism experiments are significant because they provided some of the first scientific examples of plant sensitivity. As the leading botanical taxonomist, John Ray (1628-1705) would have recognized the taxonomic implications that plant sensitivity would have on the classical distinction between plants and animals.

  19. Phototropism: the Response of Seedlings to Light

    The technique for this experiment helps students design an investigation to find out more about this tropic response. Students germinate seedlings in enclosed containers, with a coloured filter over the small hole allowing light in. Students predict which seedlings will demonstrate phototropism, and which will remain unaffected.

  20. Plant Phototropism Experiment For Students

    With each box, cut a hole 2" in diameter about 3 in. from the bottom. Take the clear cellophane and put it over the hole on one box. On the other box, place red cellophane over the hole. Now, one box will let every kind of light in and one box will only let red light in. Place your plants in the boxes, one in each.

  21. Demonstrate experiment to show A phototropism B.geotropism C

    1. Demonstration of Phototropism in Stem: Take a potted seedling of a plant. Place it in heliotropic (phototropic) chamber with a small hole on one side (). Close the top of the chamber. Place it in a position that allows sufficient light to pass through the hole. Keep it there for a couple of days. Open the chamber and observe the seedling.

  22. From "Human-Centered" to "System-Oriented": Eco ...

    The Cheonggyecheon and Yanweizhou projects aim to respond to the emerging challenges of increasing urban heat and air pollution, as well as flood hazards, respectively. In the end, several nature-based approaches and soft intervention-oriented insights are outlined for both new construction and rehabilitation of BGIs to strengthen urban resilience.