Plant Intelligence Thread

billy4479

Moderator
How smart are your plants ? Can they think feel hear /see or know if your lights on ....well i asked my self the same Q,s as i inbark on my internet gathering web knowledge mission ...I am going to post what i find on this thread as as always Id like to hear what your feelings are on the subject but id also like those of you who would like join in on the data side to to post real data findings and what everelse you have herd on the subject ..youtube clips anything id like to have this advanced forum filled with insite into every topic we can come up with ........
 

billy4479

Moderator
From Wikipedia, the free encyclopedia


here is some of the eazy to find info on them already off to a good start


In botany, plant intelligence is the ability of plants to sense the environment and adjust their morphology, physiology and phenotype accordingly.[1] Research draws on the fields of plant physiology, ecology and molecular biology.
Intelligence is an umbrella term describing abilities such as the capacities for abstract thought, understanding, communication, reasoning, learning, learning from past experiences, planning, and problem solving. Studies indicate plants are capable of problem solving and communication.
Problem solving
Plants adapt their behaviour in a variety of ways:
  • Active foraging for light and nutrients. They do this by changing their architecture, physiology and phenotype.[2][3][4]
  • Leaves and branches are positioned and oriented in response to light source.[2][5]
  • Ability to detect soil volume and adapt growth accordingly independently of nutrient availability.[6][7][8]
  • Adaptively defend against herbivores.
[edit] Communication

Plants respond to volatile signals produced by other plants.[9][10]
Mechanisms

Main articles: Signal transduction, Plant neurobiology, and Plant hormone
In plants, the mechanism responsible for adaptation is signal transduction.[11][12][13][14] Plants do not have a brain or neuronal network, but reactions within signalling pathways may provide a biochemical basis for learning and memory.[15] Controversially, the brain is used as a metaphor in plant intelligence to provide an integrated view of signalling,[16] (see plant neurobiology).
Plant cells can be electrically excitable and can display rapid electrical responses (action potentials) to environmental stimuli. These action potentials can influence processes such as actin-based cytoplasmic streaming, plant organ movements, wound responses, respiration, photosynthesis and flowering.[17][18][19][20]
Senses in plants


This section may require cleanup to meet Wikipedia's quality standards. (Consider using more specific cleanup instructions.) Please help improve this section if you can. The talk page may contain suggestions. (August 2010)
Main article: Plant perception (physiology)
Plants have many strategies to fight off pests. For example, they can produce different toxins (phytoalexins) against invaders or they can induce rapid cell death in invading cells to hinder the pests from spreading out. These strategies depend on quick and reliable recognition-systems.[edit] Smell
Wounded tomatoes are known to produce the volatile odour methyl-jasmonate as an alarm-signal.[21] Plants in the neighbourhood can then smell the danger and prepare for the attack by producing chemicals that defend against insects or attract predators.[21]
Light and electromagnetic waves

Main articles: Photomorphogenesis and photoperiodism
Many plant-organs contain photo-sensitive compounds (phototropins, cryptochromes and phytochromes) each reacting very specifically to certain wavelengths of light. These light-sensors tell the plant if it's day or night, how long the day is (photoperiodism), how much light is available and from where the light comes. Plants also can detect harmful ultraviolet B-rays and then start producing pigments which filter out these rays.[22]
Touch

Main article: Thigmotropism
The mimosa plant (Mimosa pudica) makes its thin leaves point down at the slightest touch and carnivorous plants such as the Venus flytrap snap shut by the touch of insects. But a sense of touch is something every plant has, as Coughlin describes: "Ordinary plants need a sense of touch to respond to the buffeting of the wind, which can cause damage to foliage. They try to resist wind by strengthening tissues that are being swayed. The extra energy expended stiffening tissue can cost farmers dear, however. One experiment showed that when maize plants are shaken for 30 seconds each day, yields drop by 30 to 40% compared with unshaken plants" (New Scientist).[Full citation needed]
Hearing

Main article: Thigmomorphogenesis
Mechanical perturbation can also be detected by plants.[23] Jasmonate levels also increase rapidly in response to mechanical perturbations such as tendril coiling.[24]
Mordecai Jaffe (Wake Forest University) used an instrument that made a loud "warble" and got a doubling in the growth of dwarf pea plants. Jaffe suspects that the plant hormone gibberellic acid, which is instrumental in shoot elongation and seed germination, is involved in the "hearing" response. When Jaffe added chemicals to the pea plants inhibiting the biosynthesis of this hormone, he was unable to reproduce the original effects.[citation needed]
Poplar stems can detect reorientation and inclination (equilibrioception).[25]
Criticism

It has been argued that although plants are capable of adaptation, it should not be called intelligence. "A bacterium can monitor its environment and instigate developmental processes appropriate to the prevailing circumstances, but is that intelligence? Such simple adaptation behaviour might be bacterial intelligence but is clearly not animal intelligence."[26] However, plant intelligence fits with the definition of intelligence proposed by David Stenhouse in a book he wrote about evolution where he described it as "adaptively variable behaviour during the lifetime of the individual".[27]
It is also argued that a plant cannot have goals because operational control of the plant's organs is devolved.[26]
[edit] History

Charles Darwin studied the movement of plants and in 1880 published a book The Power of Movement in Plants. In the book he concludes:
It is hardly an exaggeration to say that the tip of the radicle thus endowed [..] acts like the brain of one of the lower animals; the brain being situated within the anterior end of the body, receiving impressions from the sense-organs, and directing the several movements.
Indian scientist Sir Jagdish Chandra Bose began to conduct experiments on plants in the year 1900. He found that every plant and every part of a plant appeared to have a sensitive nervous system and responded to shock by a spasm just as an animal muscle does.[28][29]
Bose's experiments stopped at this conclusion, but American polygraph expert Cleve Backster conducted research that led him to believe that plants can communicate with other lifeforms.[30][31] Backster's interest in the subject began in February 1966, when Backster wondered if he could measure the rate at which water rises from a philodendron's root area into its leaves. Because a polygraph or "lie detector" can measure electrical resistance, and water would alter the resistance of the leaf, he decided that this was the correct instrument to use. After attaching a polygraph to one of the plant's leaves, Backster claimed that, to his immense surprise, "the tracing began to show a pattern typical of the response you get when you subject a human to emotional stimulation of short duration".
 

cannawizard

Well-Known Member
:joint:


Smarty Plants: Inside the World's Only Plant-Intelligence Lab :hump:





The "plantoid" is a concept robot for exploring Mars. Its roots would explore the soil, while power and telecommunications are provided by the main stem and the solar "leaves."

Image: Courtesy International Laboratory of Plant Neurobiology







SESTO FIORENTINO, Italy -- Professor Stefano Mancuso knows it isn't easy being green: He runs the world's only laboratory dedicated to plant intelligence.

At the International Laboratory of Plant Neurobiology (LINV), about seven miles outside Florence, Italy, Mancuso and his team of nine work to debunk the myth that plants are low-life. Research at the modern building combines physiology, ecology and molecular biology.

"If you define intelligence as the capacity to solve problems, plants have a lot to teach us," says Mancuso, dressed in harmonizing shades of his favorite color: green. "Not only are they 'smart' in how they grow, adapt and thrive, they do it without neuroses. Intelligence isn't only about having a brain."

Plants have never been given their due in the order of things; they've usually been dismissed as mere vegetables. But there's a growing body of research showing that plants have a lot to contribute in fields as disparate as robotics and telecommunications. For instance, current projects at the LINV include a plant-inspired robot in development for the European Space Agency. The "plantoid" might be used to explore the Martian soil by dropping mechanical "pods" capable of communicating with a central "stem," which would send data back to Earth.

The idea that plants are more than hanging decor at the dentist's office is not new. Charles Darwin published The Power of Movement in Plants -- on phototropism and vine behavior -- in 1880, but the concept of plant intelligence has been slow to creep into the general consciousness.

At the root of the problem: assuming that plants have, or should have, human-like feelings in order to be considered intelligent life forms, Mancuso says.







Professor Mancuso blends in with the greenery. He touches a formerly neglected office plant.

Photo: Nicole Martinelli



After the folksy 1970s hit book and stop-motion film The Secret Life of Plants, which maintained, sans serious research, that greenery had feelings and emotions, the scientific community has avoided talking about smarty plants.





So while there has been a bumper crop of studies demonstrating that green matter can be nearly as sophisticated as gray matter -- especially when it comes to signaling and response systems, few talk about intelligence.

To christen the lab in 2004, Mancuso decided to use the controversial term "plant neurobiology" to reinforce the idea that plants have biochemistry, cell biology and electrophysiology similar to the human nervous system. But although LINV is part of the University of Florence -- where Mancuso teaches horticulture -- funds for this fertile field of research weren't forthcoming.

Studies at LINV were eventually given lymph -- 1 million euro so far, with about 500,000 euro to come -- from the Ente Cassa di Risparmio di Firenze, a bank foundation that mainly supports cultural events and art restorations.

What convinced them to provide seed money?

"Looking beyond the name at the research," says Paolo Blasi, a physics professor at the university who's on LINV's board of directors. "It sounds almost like a pseudoscientific field, but now even skeptics are convinced because of the validity of the work."

In addition to studies on the effects of music on vineyards, the center's researchers have also published papers on gravity sensing, plant synapses and long-distance signal transmission in trees. One important offshoot of the research activity is an international symposium on plant neurobiology. Next year's meeting will be held in Japan.

Leopold Summerer, advanced-concepts team coordinator at the European Space Agency, remembers that the term "plant intelligence" raised a few eyebrows when collaboration with the lab was proposed -- even on a multidisciplinary think-tank team that's used to pondering ideas out of left field. Nonetheless, Summerer says plant research may provide important ideas.

"Biometrics can provide some of the most inspiring resources for us," he says. "Solutions found by nature that might not seem related to real engineering problems at first sight actually are related and give technical solutions."



Radical as the LINV sounds, if it weren't for a lone sugarcane stalk perched on a cabinet, the lab looks like any other.

While white-coated researcher Luciana Renna patiently tests for DNA markers, molecular biologist Giovanni Stefano analyzes data on two computer monitors around the corner.

During a visit to the lab's two greenhouses -- where research is being conducted on the effects of light on olive trees and reactions in Venus flytraps and the Mimosa pudica -- Mancuso points out a few neglected office plants sent there for a little TLC.

Mancuso, however, is no plant-whisperer. Under-tended plants are a long way from understanding sweet nothings spoken softly to them, he explains.

"Plants communicate via chemical substances," Mancuso says. "They have a specific and fairly extensive vocabulary to convey alarms, health and a host of other things. We just have sound waves broken down into various languages, I don't see how we could bridge the gap."
 

Joedank

Well-Known Member
Yes, playing music does have some effect on plant growth. Here is an abstract from Ultrasonics Volume 41, Issue 5, July 2003, Pages 407-411 by Yu-Chuan Qin, Won-Chu Leeb, Young-Cheol Choi and Tae-Wan Kim: "The effects of two different sonic exposures on two vegetables, namely Chinese cabbage and cucumber at two growth stages, including seedlings and mature plants were investigated. The 3 h exposures included either 20 kHz sound waves or "green music" that comprised classic music and natural sounds such as those of birds, insects, water, etc. Analysis of variance between groups (ANOVA) was used to determine the appropriate statistics parameters for the different treatments. Both exposures caused significant elevations in the level of polyamines (PAs) and increased uptake of oxygen O2 in comparison with the controls. For Chinese cabbage the highest PAs' levels were determined for both seedlings and mature plants that were exposed to "green music". The oxygen uptake in Chinese cabbage also increased as a result of sonic exposures, and the highest oxygen uptake was also observed after "green music" treatment. For cucumber, the highest content of PAs for both seedlings and mature cucumber plants was determined as a result of 20 kHz ultrasound exposure. 20 kHz exposure of mature plants also resulted in the highest level of oxygen uptake. No statistically significant differences in the vitamin C level were determined between the different sonic treatments and sham exposed vegetables. "
My plants hear music 8 hrs a day especial if foliar feeding
 

cannawizard

Well-Known Member
--from the .pdf link, this caught my attention..

23.7

Colour Signals

Colour change and colour pattern are powerful tools in plant–animal communication.

The functional and evolutionary importance of colour signalling

in animals has received great consideration in zoology, resulting in

numerous theories and wide experimentation (Majerus 1998). In contrast,

with the exclusion of studies on the colour importance for the attraction

of pollinators to flowers (Chittka et al. 1999) and frugivors to fruit (Ridley

1930), the biological relevance of colour has been extensively underestimated

in plant sciences. Yet, visual signals sent to animals are usuallymore

efficient than olfactory signals on long-distance signalling, owing to the

great influence of the environment on the diffusion of volatiles (Dobson

1994; Anderson and Dobson 2003).

One the most exciting colour signals produced by plants is the bright autumncolorationdisplayedbymanydeciduous

trees.Why some tree species

make this spectacular exhibition of colour is one of the most puzzling

questions in tree biology. The usual explanation is that autumn colours are

simply a secondary and mere side effect of leaf senescence. In autumn the

degeneration of chloroplasts and the degradation of chlorophyll pigments

in colourless low molecular products allows the red and yellow pigments

(carotenoids and flavonoids) to appear fromthe background (Sanger 1971;

Goodwin and Mercer 1983). This point of view, however, overlooks two

important facts: many trees do not show any bright colouration in autumn

and, more important, there are numerous pieces of evidence that colour

change is also due to the synthesis of new pigments (Chang et al. 1989;

Matile et al. 1992).

Two recent papers have challenged this interpretation by suggesting that

these red and yellow leaf colours are an honest signal of tree’s ability to

defend itself against potential insect pests (Archetti 2000; Hamilton and

Brown 2001). Hamilton and Brown’s theory explains that the bright colour

of autumn foliage is not just a side effect of chlorophyll reabsorption but

acts as a signal, for aphids that are looking for places to lay their eggs, to

indicate that the tree has invested heavily in chemical defence, and it is,

therefore, not suitable for aphids. Hamilton and Brown (2001) predicted

 

billy4479

Moderator
Electric signalling in fruit trees in response to water applications and light–darkness conditions



Summary

A fundamental property of all living organisms is the generation and conduction of electrochemical impulses throughout their different tissues and organs, resulting from abiotic and biotic changes in environmental conditions. In plants and animals, signal transmission can occur over long and short distances, and it can correspond to intra- and inter-cellular communication mechanisms that determine the physiological behaviour of the organism. Rapid plant and animal responses to environmental changes are associated with electrical excitability and signalling. The same molecules and pathways are used to drive physiological responses, which are characterized by movement (physical displacement) in animals and by continuous growth in plants. In the field of environmental plant electrophysiology, automatic and continuous measurements of electrical potential differences (ΔEP) between plant tissues can be effectively used to study information transport mechanisms and physiological responses that result from external stimuli on plants. A critical mass of data on electrical behaviour in higher plants has accumulated in the last 5 years, establishing plant neurobiology as the most recent discipline of plant science. In this work, electrical potential differences were monitored continuously using Ag/AgCl microelectrodes, which were inserted 15 mm deep into sapwood at various positions in the trunks of several fruit-bearing trees. Electrodes were referenced to an unpolarisable Ag/AgCl microelectrode, which was installed 5 cm deep in the soil. Systematic patterns of ΔEP during day–night cycles and at different conditions of soil water availability are discussed as alternative tools to assess early plant stress conditions. This research relates to the adaptive response of trees to soil water availability and light–darkness cycles
 

billy4479

Moderator
So do you think its possible if your plants had turned yellow late in flowering that if aphids did come the might go for your greener plants first ?
 

cannawizard

Well-Known Member
http://altweb.astate.edu/electronicjournal/suza_et_al.htm

CLICK HERE FOR DMT-Plant sourced communication pdf

View attachment 1789383
Root–microbe communication that can lead to defense



responses



Root exudates also act as antimicrobials against rhizospheric



microflora, providing the plant with defensive



advantages. Collectively, plants produce a compositionally



diverse array of




.100 000 different low-molecular-mass



natural products, known as secondary metabolites




[23].



The rich diversity of secondary metabolites arises partly



because of selection for improved defense mechanisms



against a broad array of microbes, insects and plants.



Although such diversity has made it difficult to apply



conventional molecular and genetic techniques to determine



the functions of natural products in plant defense,



examining gene expression in model plants and microorganisms



should lead to a better understanding of



the processes mediating
 

billy4479

Moderator
Characteristics of Electrical Signals in Poplar and Responses in Photosynthesis1

Silke Lautner, Thorsten Erhard Edgar Grams, Rainer Matyssek and Jörg Fromm*
Fachgebiet Angewandte Holzbiologie, Technische Universität München, 80797 Munich, Germany (S.L., J.F.); and Lehrstuhl Ökophysiologie der Pflanze, Technische Universität München, 85354 Freising, Germany (T.E.E.G., R.M.)

ABSTRACT


To gain an understanding of the role of electrical signaling in trees, poplar (Populus trichocarpa, Populus tremula x P. tremuloides) shoots were stimulated by chilling as well as flaming. Two kinds of signal propagation were detected by microelectrode measurements (aphid technique) in the phloem of leaf veins: (1) basipetal, short-distance signaling that led to rapid membrane hyperpolarization caused by K+-efflux within the leaf lamina; and (2) acropetal, long-distance signaling that triggered depolarization of the membrane potential in the leaf phloem. In the latter, the depolarizing signals travel across the stem from the manipulated leaves to adjacent leaves where the net CO2 uptake rate is temporarily depressed toward compensation. With regard to photosystem II, both heat-induced long-distance and short-distance signaling were investigated using two-dimensional "imaging" analysis of chlorophyll fluorescence. Both types of signaling significantly reduced the quantum yield of electron transport through photosystem II. Imaging analysis revealed that the signal that causes yield reduction spreads through the leaf lamina. Coldblocking of the stem proved that the electrical signal transmission via the phloem becomes disrupted, causing the leaf gas exchange to remain unaffected. Calcium-deficient trees showed a marked contrast inasmuch as the amplitude of the electrical signal was distinctly reduced, concomitant with the absence of a significant response in leaf gas exchange upon flame wounding. In summary, the above results led us to conclude that calcium as well as potassium is involved in the propagation of phloem-transmitted electrical signals that evoke specific responses in the photosynthesis of leaves.

Electrical signaling in plants was first revealed in the 1870s in insectivorous plants by Burdon-Sanderson (1873) and Darwin (1875). In the 20th century, evidence for the existence of action potentials was presented in a broad array of plant species, irrespective of the presence of rapid leaf movements (Bose, 1924; Pickard, 1973). Most of the research on electrical signaling dealt with responses evoked by wounding of aboveground organs, providing insights into various processes of plant physiology. Molecular tools made it possible to detect rapid changes in gene expression (Davies and Schuster, 1981; Stankovic and Davies, 1997) as well as activation of proteinase inhibitor genes within plants (Bowles, 1990; Ryan, 1990; Wildon et al., 1992) upon wounding, even across long distances. Wildon et al. (1992) showed the chemical signals evoked by wounding in the phloem to be significantly slower than the rapid changes in membrane potential. Electrical signals that were generated and transmitted from distant plant parts arrived at responding tissues well before the initiation of transcript accumulation. Vian et al. (1999) induced rapid and systemic accumulation of chloroplast mRNA-binding protein transcripts, in tomato (Lycopersicon esculentum), after flame stimulus. In addition to translation and transcription, evidence exists for a role of electrical signals in many processes of plant life, including respiration (Dziubinska et al., 1989; Filek and Koscielniak, 1997), water uptake (Davies et al., 1991), phloem unloading (Fromm, 1991), and phloem translocation (Fromm and Bauer, 1994) as well as fertilization (Fromm et al., 1995). Recently, this account was extended by a study on the inhibition of photosynthesis in Mimosa pudica (Koziolek et al., 2004) upon flame wounding, which demonstrated that electrical signals triggered transient changes in chlorophyll fluorescence (PSII electron quantum yield) and leaf gas exchange. Moreover, in this latter case, the transport of chemicals in the phloem was far too slow to account for the induced changes.
In trees, electrically induced action potentials were measured in willow (Salix viminalis) shoots, showing that calcium influx as well as potassium and chloride efflux are involved in the propagation of signals within the phloem (Fromm and Spanswick, 1993). Using a vibrating electrode in combination with the standard microelectrode technique, an apparent efflux of anions and cations of 200 to 700 pmol cm–2 per action potential was assessed in willow roots (Fromm et al., 1997). These changes in the cellular ion concentrations may be important in intracellular signaling, whereas communication over long distances in trees may be achieved through phloem-transmitted electrical signals. Bridging long distances, these rapid signals possess the capacity for coordinating physiological activities in trees. To ascertain potential functions of long-distance electrical signals in trees that may have a bearing on photosynthesis, noninvasive techniques were employed in this study on poplar (Populus trichocarpa, Populus tremula x P. tremuloides) trees about 80 cm tall. After flame wounding, electrical signals were measured in the phloem by means of the aphid technique (Fromm and Eschrich, 1989). Care was taken to avoid any kind of puncturing stress in the trees, which were grown under nonlimiting nutrient supply or under conditions of calcium deficiency. In parallel, leaf responses were measured in gas exchange (via porometry) and chlorophyll fluorescence (by a two-dimensional imaging approach; Koziolek et al., 2004) at various distances from the wounding site. The hypothesis was tested that (1) heat-induced long-distance electrical signaling affects photosynthesis, and (2) calcium deficiency reduces the capacity for signal transmission. Since intracellular calcium is one of the major elements in the signal transduction pathways of plant cells (Okihara et al., 1991; Trewavas, 2000), calcium-deficient trees were expected to exhibit a difference in behavior.

RESULTS



Long-Distance Electrical Signaling
To detect electrical signals in poplar shoots, the membrane potential was measured either in the leaf mesophyll or in the phloem via severed aphid stylets, at a point in the upper stem (Fig. 1, electrode B) or at the first mature leaf (electrode A). The resting potential of the measured phloem cells ranged between –116 mV to –165 mV in 10 experiments with different plants and was similar to the sieve tube potentials in M. pudica (Fromm, 1991) and maize (Zea mays; Fromm and Bauer, 1994). In agreement with the common observation that electric activity in plants can be incited through contact with ice water, such stimulation of the tip of leaf 1 resulted, in the sieve tubes, in a basipetally propagating electrical signal with a hyperpolarizing amplitude of 25 mV (recorded at either microelectrode position; Fig. 2, top graph). Transmission velocity was 4 to 8 mm s–1. By contrast, when stimulating the lower part of the stem with ice water at the position of leaf 4, action potentials with depolarizing amplitudes of 12 to 20 mV and velocities similar to those reported above were recorded at the two microelectrode positions (Fig. 2, bottom graph). The original resting potential was reestablished about 3 to 4 min after depolarization, thus indicating typical criteria inherent to action potentials.


View larger version (57K):
In this window
In a new window

Figure 1. Experimental arrangement of electrical potential recordings. A, The plant was heat stimulated for 3 s by the flame of a lighter either at the tip of leaf 1 or at the base of leaf 4. Cold stimulation (chilling) was applied to the tip of leaf 1 or the lower stem. To disrupt the propagation of electrical signals, a coldblock of 4°C was applied to the stem between the second and third mature leaf underneath the apex. Coldblocking is not equivalent to cold stimulation (chilling) where ice water is rapidly applied for a few seconds, whereas the coldblock at the stem remains in situ permanently in order to block the excitability of the plasma membranes. B, top, An aphid sucking on the underside of a leaf, in the phloem of the primary vein. B, bottom, After the aphid had been severed from its mouthparts by a microscope laser, the stylet stump exuded sieve tube sap to which the tip of a microelectrode was attached by using a micromanipulator.




View larger version (19K):
In this window
In a new window

Figure 2. Electrical signals following cold stimulation of detached poplar shoots. As described in Figure 1, the plant was cold stimulated by ice water (4°C) for several seconds at the tip of leaf 1 or at the mid of the stem. Electrical signals were monitored in leaf 1 (electrode A) or in the upper stem (electrode B). The arrows denote the instant of stimulation. Upward and downward deflections signify depolarization and hyperpolarization of the membrane potential respectively. The curves are typical examples from five measurements with different plants.


Flame stimulation of the tip of leaf 1 evoked a propagating electrical signal with a hyperpolarizing amplitude of approximately 25 mV as measured in the phloem by electrodes A and B as well as in the mesophyll of the flame-wounded leaf, after electrode A had been inserted into the leaf mesophyll (Fig. 3, top graph). Signal transmission velocity in basipetal direction was 1 to 2 mm s–1. Following flame stimulation of the base of leaf 4, either electrode recorded an irregularly shaped propagating electrical signal with an amplitude of over +50 mV in the phloem as well as in the mesophyll of leaf 1 (Fig. 3, central graph). In contrast to the action potentials evoked by stem chilling (Fig. 2, bottom graph), the flame-induced signals reflected failure in reestablishing the original resting potential, indicating that the voltage changes are different from action potentials. The transmission velocity was 1 to 2 mm s–1 in acropetal direction and, hence, similar to that in basipetally propagating flame-induced signals. To disrupt the propagation of electrical signals, a coldblock of 4°C was applied to the stem between the second and third mature leaf underneath the apex. In contrast to the unchilled plants, the resting potential of the phloem depolarized only slightly at either electrode (Fig. 3, bottom graph), indicating a strong decrease in signal amplitudes.


View larger version (22K):
In this window
In a new window

Figure 3. Responses of the membrane potential of the phloem and the mesophyll to flame stimulation of leaf 1 and leaf 4. Stimulation of the tip of leaf 1 evoked the propagation of electrical signals of a hyperpolarized nature (top graph), while stimulation of leaf 4 generated signals of a depolarized nature (central graph). Applying a coldblock to the stem between leaves 2 and 3 significantly reduced the amplitude of the transmitted signal (bottom graph). The curves are typical examples from five measurements with different plants.


It is remarkable that, independent of the type of stimulation (chilling, heating), signals moving in basipetal direction change toward the negative direction, which contrasts with the opposite response in acropetally traveling signals. The Ca2+-influx/Cl–-efflux/K+-efflux sequence in signal generation may explain the depolarizing direction, but different fluxes are needed to explain signals of the negative sign, such as K+-efflux preceding Cl–-efflux. The latter was proved by blocking K+ channels with 1 mM tetraethylammoniumchloride (TEA+), acting on both sides of the membranes (Wong and Adler, 1986) and applied to the artificial pond water (APW) medium of the stem (Fig. 1). After 1 d, the inhibitor had been distributed within the plant. When these plants were flame stimulated at the tip of leaf 1, no hyperpolarizing signal was measured in the phloem (Fig. 4, top graph). Obviously, transient hyperpolarization is caused by the transient opening of K+ channels in the plasma membrane.


View larger version (27K):
In this window
In a new window

Figure 4. Typical responses of the membrane potential in phloem cells upon flame stimulation in TEA+-treated as well as calcium-deficient plants. Upon application of TEA+, a K+ channel blocker, flame stimulation of leaf 1 did not evoke an electrical signal (top graph). Also in Ca-deficient plants, stimulation of leaf 1 failed to generate any signal (central graph), and wounding of leaf 4 caused weak changes in the membrane potential (bottom graph).


As the ionic mechanism of excitation is based on the movement of potassium, chloride, and calcium (Tazawa et al., 1987; Fromm and Spanswick, 1993), the objective of experiments involving the stimulation of calcium-deficient plants was to characterize the excitation mechanism in poplar. The Ca-deficient plants were distinctly smaller and their leaf size was reduced compared with that in nonlimited plants. Flame wounding of the tip of leaf 1 evoked no detectable electrical signals in the phloem at either electrode position (Fig. 4, central graph). However, signals with strongly reduced amplitudes were measured in acropetal direction after flame stimulation of leaf 4 (Fig. 4, bottom graph), perhaps reflecting the physiological significance of the upwards transmission of electrical signals in plant life.

Chlorophyll Fluorescence Imaging
No photosynthetic response was found upon stimulation of leaves or stem with ice water. However, flaming of the tip of leaf 1 caused a substantial decrease in the electron quantum yield of PSII (Fig. 5A). This effect occurred 80 s after flame stimulation in the intercostal regions of the central part of the lamina, at a distance of 3 cm from the leaf tip. The inhibitory response spread basipetally throughout the leaf, showing a delay in the arrival of the electrical signal at electrode B (approximately 60 s after flame wounding of the leaf tip) in relation to electrode A (after only 2 s; Fig. 3, top graph). In the flame-stimulated leaf, the fluorescence response reached a minimum after about 300 s, prior to incipient recovery.



View larger version (216K):
In this window
In a new window

Figure 5. Spatiotemporal changes of PSII electron quantum yield, ΔF/F'm assessed by chlorophyll fluorescence imaging. The imaged area (l, 22 mm; w, 17 mm) covers the center of leaf 1. A, The tip of leaf 1 was flame stimulated at a distance of 3 cm. Times are given in relation to the instant of injury (at time 0). Changes in ΔF/F'm took 80 s to become apparent. A false-color shift from blue to yellow in the intervein area (equivalent to a lowering of PS electron quantum yield from 0.6 to about 0.2) indicates the decrease in the electron quantum yield of PSII. B, Leaf 4 was wounded at a distance of over 11 cm from the imaged area. Changes in fluorescence were not visible before 240 s had elapsed. After about 420 and 600 s, the left and right side of the leaf, respectively, reached a minimum in the fluorescence response (white and black arrow, respectively). The bar translates the false color code into values of ΔF/F'm.


Flaming of leaf 4 caused a decrease in the electron quantum yield of PSII in leaf 1, evidence of which was found in the leaf veins 240 s of the latter after flame stimulation, and indicating that the signal spreads via the veins into the mesophyll (arrival after 300 s; Fig. 5B). The inhibitory response propagated acropetally throughout the leaf and was delayed in relation to the electrical signal that arrived in leaf 1 only 80 s after flaming of leaf 4 (Fig. 3, central graph). A minimum in the time course of the fluorescence response was reached in the left half of the lamina of leaf 1 after about 420 s (Fig. 5B, white arrow) prior to incipient recovery. However, it took about 600 s in the right half of the leaf for the minimum to occur (black arrow).
The highly resolved time course of the fluorescence response shows the decrease in PSII quantum yield to occur simultaneously in both the veins and the intercostal regions of leaf 1 taking 160 s from the time of flame stimulation of the leaf tip (distance approximately 5 cm; Fig. 6A). By contrast, a distinct two-step response was observed in leaf 1 during acropetal signaling when leaf 4 was flame stimulated, as the response in veins was faster and stronger than in the intercostal regions (Fig. 6B). In these latter areas, the transient response was delayed by about 60 s compared with the occurrence in the veins. The differences in the response between long-distance signaling along the shoot (Fig. 6B) and short-distance signaling within the same leaf (Fig. 6A) reveal, in the first case, that the signals spread into the mesophyll via the veins, whereas in the latter case, signals travel across veins and mesophyll at similar velocities.


View larger version (52K):
In this window
In a new window

Figure 6. Kinetic changes of PSII electron quantum yield, ΔF/F'm, measured by fluorescence imaging at representative sites within leaf 1 after flame wounding of A, its leaf tip, and B, the base of leaf 4 at time 0. Blue and red symbols represent vein and intervein field areas, respectively. Data shown are means ± SD, n = 5. Parts of imaged areas showing representative vein and intervein field areas used for kinetic analysis are shown below the graphs.



Leaf Gas Exchange
In leaf 1, the net CO2 uptake rate (JCO2) sharply dropped to about compensation 30 s after flame stimulation of the leaf tip and stayed there for about 90 s before subsequent recovery (Fig. 7A). At the same time, the stomatal conductance (gH2O) remained stable, indicating the absence of stomatal movements. The finding of a photosynthetic reaction not showing any reaction in the gH2O is astonishing. However, the internal CO2 concentration increased transiently (data not shown), indicating that the relationship between JCO2 and gH2O is correct. A similar response occurred in leaf 1, when leaf 4 was flame stimulated at a distance of over 10 cm. At 120 s after stimulation, the JCO2 decreased immediately to compensation, staying there for 100 s and then recovering almost completely (Fig. 7B). Chilling the stem slowly down to a temperature of +4°C (coldblock) between leaves 2 and 3 showed that the gas exchange of leaf 1 remained unaltered (Fig. 7C), concomitant with the suppression of the long-distance electrical signal (Fig. 3, bottom graph). This coincidence indicates that the electrical signal has a modifying impact on CO2 uptake. Experimentation with calcium-deficient plants did not result in any significant response in gas exchange after flame stimulation of leaf 4 (Fig. 7D), which is consistent with the distinctly reduced amplitude of the recorded electrical signal (Fig. 4, bottom graph).


View larger version (21K):
In this window
In a new window

Figure 7. Typical responses of JCO2 and gH2O of leaf 1 upon heat stimulation of its tip (A) or of leaf 4 (B). The arrows denote the instant of injury. For heat stimulation of the leaf tip, mean of JCO2 before stimulation was 1.35 ± 0.15 and mean of JCO2 at transient minimum was –0.2 ± 0.2, while mean of gH2O was 105 ± 15, n = 5. For heat stimulation of leaf 4, mean of JCO2 before stimulation was 1.38 ± 0.21 and mean of JCO2 at transient minimum was 0.48 ± 0.5, while mean of gH2O was 108 ± 17, n = 5. C, Chilling the stem with ice water evoked no response in gas exchange. D, No significant response in gas exchange was recorded in calcium-deficient plants following flame stimulation of leaf 4.


To determine the velocity of chemical signaling, the petioles of excised leaves were fed with 14C-labeled Suc. After a 600-s translocation, macroautoradiography showed that the primary vein became labeled and 14C-Suc extended from the vein across the entire lamina (data not shown). Since the velocity of electrical signal transmission following flame stimulation was 1 to 2 mm s–1 in both acropetal and basipetal direction, autoradiography proved chemical signaling to be much too slow to account for the photosynthetic response after flame stimulation
DISCUSSION


The results presented here show that electrical signals propagated over long distances as well as short distances are capable of modifying photosynthesis in trees. Previous studies had reported on the capacity of many plant species to generate and transmit action potentials as well as variation potentials (Pickard, 1973; Davies, 1987). In ecophysiology, electrical signals play a key role in the communication of environmental stimuli within plants, apparently acting within and across cells, tissues, and organs (Volkov and Mwesigwa, 2001). For instance, the pesticide 2,4-dinitrophenol, considered an environmental problem in agriculture as well as a human health hazard, induced fast action potentials and decreased the variation potential in soybean (Glycine max; Mwesigwa et al., 2000). Pentachlorophenol too, a known pollutant and uncoupler of oxidative phosphorylation, induced action potentials in soybean (Volkov et al., 2000). Recent experiments on signal transmission in M. pudica demonstrated that a link exists between flame-induced electrical signals and photosynthetic responses, as inferred from two-dimensional imaging analysis of chlorophyll fluorescence in combination with gas exchange assessment (Koziolek et al., 2004).
This study on poplar confirms the latter findings for trees. Moreover, our results demonstrate that different stimulation types and positions incite characteristic electrical signals, each with a specific influence on photosynthesis. We used the aphid technique to measure electrical signals in phloem cells that share fundamental properties with nerve cells in animals, i.e. the existence of excitable membranes by which electrical excitations can be transmitted from cell to cell. In poplar, the nature of the signal depends on the traveling direction. Basipetally propagating signals (induced by chilling as well as flaming) showed negative voltage changes, whereas acropetally moving signals were characterized by transient membrane depolarization (Figs. 2 and 3). The ionic mechanism of action potentials in plants, i.e. depolarizing signals with positive sign, is based on calcium influx as well as initial chloride efflux followed by potassium efflux (Beilby and Coster, 1979; Kikuyama and Tazawa, 1983; Lunevsky et al., 1983; Tsutsui et al., 1986; Kikuyama, 1987; Fromm and Spanswick, 1993). To explain signals of the negative sign, we blocked K+ channels by TEA+ and found that no basipetally transmitted signals were induced after flame stimulating the tip of leaf 1 (Fig. 4, top graph), indicating that K+ efflux causes the hyperpolarization of basipetally transmitted signals. A similar mechanism is reported for the unicellular green alga Eremosphaera viridis, where a sudden blockage of photosynthetic electron transport by darkening causes a transient hyperpolarization of the plasma membrane (Schönknecht et al., 1998). In Eremosphaera, the transient hyperpolarization is due to the opening of K+ channels that is caused by a rapid transient elevation of the cytosolic free calcium concentration. In poplar plants grown under calcium deficiency, phloem cell excitability was completely inhibited after flaming of the leaf tip (compare with Fig. 4). This result indicates first, that calcium channels are involved in the induction of basipetally transmitted electrical signals, confirming the observation that calcium channels and transporters form the basis of a complex calcium signaling network (Trewavas and Malho, 1997; Trewavas, 2000). Second, in optimum nutrient-supplied plants, blockage of the signal by TEA+ indicates that K+ efflux is responsible for the occurrence of hyperpolarization, most likely induced by calcium as reported for Eremosphaera (Schönknecht et al., 1998).
In the poplars of this study, electrical signals were induced by chilling as well as by flame wounding. As regards chilling, amplitudes and dynamics of the acropetally transmitted signals were typical of plant action potentials, as previously demonstrated in maize (Fromm and Fei, 1998) and, for trees, in willow (Fromm and Spanswick, 1993). Action potentials are generally self-amplifying signals propagated at constant velocity. It is assumed that they are dependent on voltage-gated ion channels and are capable of propagating through any living cells sharing common membranes, but they are most evident in sieve tubes. In contrast to action potentials, flame stimulation causes the appearance of a slowly moving, irregular, so-called variation potential (Sibaoka, 1966; Sibaoka, 1969). There is widespread agreement in the literature that the variation potential is not a self-propagating signal but, rather, a local electrical response to the underlying passage of chemical substances released from the wound site and transmitted through the xylem by hydraulic dispersal. Malone (1992) reported on hydraulic signals in wheat leaves, traveling at a velocity of at least 100 mm s–1 from the site of stimulation. However, using a combination of electrometer and laser-interferometer, Tinz-Füchtmeier and Gradmann (1990) were unable to confirm the hydraulic conductance of excitation in Mimosa.
In poplar, hydraulic signals may play a role in the generation of electrical excitation. Since we managed to measure flame-induced signals in the phloem, the question arises whether pressure or chemical changes in the xylem can activate ion channels in the phloem, making it appear as if a hydraulically induced variation potential were passing through the phloem. However, as the electric signal transmission was disrupted after applying a coldblock to the stem and gas exchange did not respond to flame wounding, it is thought unlikely that hydraulic events play any role in long-distance signaling in poplar. Moreover, variation potentials certainly depend on the prevailing water status of the plant. When shoot water status is saturated, as in the case of the well-watered plants used in this study, xylem tension becomes negligible and variation potentials should not travel at all. We are therefore convinced that the phloem-transmitted flame-induced signals in poplar are self-propagating signals, independent of chemicals traveling through the xylem.
Concerning chemical signaling in the phloem that spreads from the stimulation site through the sieve tubes, the thought cannot be dismissed that photosynthesis is affected by chemical signals. Canny (1975) reported transport velocities in the phloem to range between 50 and 100 cm h–1, i.e. too slow to play a key role in the studied poplar trees. The first response in gas exchange upon wounding of the leaf tip took 30 s to travel a distance of 4 cm, whereas a chemical signal would have moved across a distance of no more than 0.4 to 0.8 cm in 30 s. In consistency with these findings, autoradiography confirmed that chemical signaling is too slow to account for the photosynthetic response (data not shown).
With regard to the photosynthetic response, only flame-evoked signals caused photosynthetic changes, whereas chill-induced signals had no impact on photosynthesis. Interestingly, the noninvasive imaging analysis of chlorophyll fluorescence revealed that short-distance signals within leaves after flame stimulation at the tip cause a simultaneous decrease in electron quantum yield of PSII in both veins and intercostal regions (compare with Fig. 6A). By contrast, long-distance signals, again generated by flame wounding of leaves but arriving after having traveled across the plant, reduce the electron quantum yield of PSII in the veins first and in the intercostal regions afterward (Fig. 6B), hence suggesting that the signal spreads via the veins into the mesophyll. Reduction in photosynthesis upon impact by electrical signals is also known for M. pudica (Koziolek et al., 2004) and tomato plants (Herde et al., 1999). Since the PSII quantum yield decreased as well as the JCO2, it can be excluded that the latter is only based on increased respiration. However, the mechanism underlying photosynthetic limitation upon impact by electrical signals requires further clarification. Subcellular alterations in ion fluxes may be involved in the photosynthetic response so that attention should focus on the translocation of ions in mesophyll cells and their chloroplasts. It is reported that enzymes in the cell wall, the plasmalemma, and the cytoplasm can show modified activities during local changes in ion concentrations (Davies, 1987). Bulychev et al. (1986, 1987) reported changes in the chlorophyll fluorescence of intact chloroplasts induced by shifts in membrane potential. The rapid impact on chlorophyll fluorescence and CO2 uptake in poplar is likely, therefore, to be caused by local changes in ion concentrations. For spinach, evidence is presented for a direct involvement of calcium in O2 formation of PSII (Vrettos et al., 2001), indicating a possible role for calcium in the photosynthetic response. In summary, this study provides evidence that rapidly evoked and phloem-transmitted electrical signals can affect photosynthesis over long distances in trees. However, the intracellular controlling points in the leaf mesophyll will be the subject of further investigations.

MATERIALS AND METHODS



Plant Material and Growth Conditions
One-year-old plants of Populus trichocarpa cv Trichobel were grown in the summer from cuttings under standard greenhouse conditions, at 25°C and 85% relative humidity, in sandy culture medium. In addition, 8-week-old plants of Populus tremula x P. tremuloides Michx. were grown hydroponically but under otherwise identical greenhouse conditions. The latter plants were provided with macro- and micronutrients in a modified Hoagland solution (Hoagland and Arnon, 1950) containing either low Ca2+ (0.1 mM) or full-strength Ca2+ (5 mM) concentration. Experiments were performed using fully developed shoots from both groups of plants. Effects of calcium deficiency on electrical signaling were studied using the hydroponically grown plants (P. tremula x P. tremuloides Michx.); all other experiments were performed on P. trichocarpa cv Trichobel.

Electrical Potential Measurements
When about six leaves had developed, plants were cut from their roots and transferred to a Faraday cage for experimentation. They had to be cut from their roots in order to detect the membrane potential by measurements from two sides of the membrane. Microelectrode tips were either inserted into the leaf mesophyll or brought into contact with the exudate droplet on a severed aphid stylet. In the latter case, contact was made at two different positions, either at some point in the upper stem (electrode B) or at the lower side of the first mature leaf underneath the apex (electrode A, Fig. 1). The reference electrode (Ag/AgCl) was immersed into the APW where the cut cross section of the excised stem had also been submerged. The APW was composed of 1.0 mM NaCl, 0.1 mM KCl, 0.1 mM CaCl2, and 1.0 mM MES, adjusted with Tris to a pH value of 6.0.
The electrical potential of the phloem was measured via severed aphid stylets. This involved introducing aphids to a leaf or the stem and allowing them to settle overnight. On the following day, they were severed from their stylets by shots from a laser beam generator (Beck, Neu-Isenburg, Germany), connected to a Zeiss microscope. Electric potential changes were measured through glass microelectrodes with tip diameters of less than 1 µm, back-filled with 3 M KCl. The microelectrode was clamped in an Ag/AgCl pellet holder (WPI) and connected to a microelectrode preamplifier (input impedance >1012 ohms) to which a WPI amplifier (model 750, WPI, Sarasota, FL) was attached. The response time of the microelectrodes is about 1 s, which is fast enough to measure electrical responses induced by heat and cold stimulation. The electrodes were inserted into the leaf mesophyll or attached to the stylet stump by micromanipulators: electrode A in leaf 1 at a mean distance of 30 to 40 mm to the stimulated tip; electrode B at some point in the stem (Fig. 1). The resistance for an electric current inside the aphid stylet is relatively low (around 109 Ω according to Wright and Fisher, 1981) compared to the high input impedance of the electric equipment used. Recordings were made between the microelectrodes and a reference electrode (in APW and connected to the cut end of the shoot). Measurement results were logged in parallel by chart recorder and computer.

Chlorophyll Fluorescence Imaging
The two-dimensional imaging approach described by Koziolek et al. (2004; Imaging-PAM Chlorophyll Fluorometer, Heinz Walz GmbH, Effeltrich, Germany) was employed to assess the spatiotemporal variations of the quantum yield of energy conversion in PSII (Siebke and Weis, 1995; Rascher et al., 2001). This method allows noninvasive determination of PSII quantum yield by the saturation pulse method (Schreiber et al., 1986; Genty et al., 1989). Blue light (470 nm) is applied to act as pulse-modulated measuring light, actinic illumination, and saturation pulses. The imaged area of leaf 1 was adapted for at least 10 min to a PPFD of 100 µmol m–2 s–1 prior to flame wounding of the leaf tip (same leaf) or an adjacent leaf. Heat treatment involved a flame stimulus lasting 3 s. Saturation pulses were given every 10 s to determine the images of fluorescence yield, F, maximal fluorescence yield, Fm', and PSII quantum yield, ΔF/Fm' = (Fm'– F)/Fm' (for nomenclature, see van Kooten and Snel, 1990).

Leaf Gas Exchange
The gas exchange in the attached leaves was measured using a steady-state CO2/water diffusion porometer (CQP130, Walz, Effeltrich, Germany) at ambient CO2 concentration of about 360 µL L–1 and a relative humidity of approx. 60%, leaf temperature of about 27°C, and PPFD of about 100 µmol m–2 s–1. Leaf 1 was attached to the porometer and its leaf tip or leaf 4 (compare with Fig. 1) were stimulated by flaming. Leaf net-CO2 exchange was calculated based on single leaf area and expressed as JCO2.
 

billy4479

Moderator
Characteristics of Electrical Signals in Poplar and Responses in Photosynthesis1

Silke Lautner, Thorsten Erhard Edgar Grams, Rainer Matyssek and Jörg Fromm*
Fachgebiet Angewandte Holzbiologie, Technische Universität München, 80797 Munich, Germany (S.L., J.F.); and Lehrstuhl Ökophysiologie der Pflanze, Technische Universität München, 85354 Freising, Germany (T.E.E.G., R.M.)

ABSTRACT


To gain an understanding of the role of electrical signaling in trees, poplar (Populus trichocarpa, Populus tremula x P. tremuloides) shoots were stimulated by chilling as well as flaming. Two kinds of signal propagation were detected by microelectrode measurements (aphid technique) in the phloem of leaf veins: (1) basipetal, short-distance signaling that led to rapid membrane hyperpolarization caused by K+-efflux within the leaf lamina; and (2) acropetal, long-distance signaling that triggered depolarization of the membrane potential in the leaf phloem. In the latter, the depolarizing signals travel across the stem from the manipulated leaves to adjacent leaves where the net CO2 uptake rate is temporarily depressed toward compensation. With regard to photosystem II, both heat-induced long-distance and short-distance signaling were investigated using two-dimensional "imaging" analysis of chlorophyll fluorescence. Both types of signaling significantly reduced the quantum yield of electron transport through photosystem II. Imaging analysis revealed that the signal that causes yield reduction spreads through the leaf lamina. Coldblocking of the stem proved that the electrical signal transmission via the phloem becomes disrupted, causing the leaf gas exchange to remain unaffected. Calcium-deficient trees showed a marked contrast inasmuch as the amplitude of the electrical signal was distinctly reduced, concomitant with the absence of a significant response in leaf gas exchange upon flame wounding. In summary, the above results led us to conclude that calcium as well as potassium is involved in the propagation of phloem-transmitted electrical signals that evoke specific responses in the photosynthesis of leaves.

Electrical signaling in plants was first revealed in the 1870s in insectivorous plants by Burdon-Sanderson (1873) and Darwin (1875). In the 20th century, evidence for the existence of action potentials was presented in a broad array of plant species, irrespective of the presence of rapid leaf movements (Bose, 1924; Pickard, 1973). Most of the research on electrical signaling dealt with responses evoked by wounding of aboveground organs, providing insights into various processes of plant physiology. Molecular tools made it possible to detect rapid changes in gene expression (Davies and Schuster, 1981; Stankovic and Davies, 1997) as well as activation of proteinase inhibitor genes within plants (Bowles, 1990; Ryan, 1990; Wildon et al., 1992) upon wounding, even across long distances. Wildon et al. (1992) showed the chemical signals evoked by wounding in the phloem to be significantly slower than the rapid changes in membrane potential. Electrical signals that were generated and transmitted from distant plant parts arrived at responding tissues well before the initiation of transcript accumulation. Vian et al. (1999) induced rapid and systemic accumulation of chloroplast mRNA-binding protein transcripts, in tomato (Lycopersicon esculentum), after flame stimulus. In addition to translation and transcription, evidence exists for a role of electrical signals in many processes of plant life, including respiration (Dziubinska et al., 1989; Filek and Koscielniak, 1997), water uptake (Davies et al., 1991), phloem unloading (Fromm, 1991), and phloem translocation (Fromm and Bauer, 1994) as well as fertilization (Fromm et al., 1995). Recently, this account was extended by a study on the inhibition of photosynthesis in Mimosa pudica (Koziolek et al., 2004) upon flame wounding, which demonstrated that electrical signals triggered transient changes in chlorophyll fluorescence (PSII electron quantum yield) and leaf gas exchange. Moreover, in this latter case, the transport of chemicals in the phloem was far too slow to account for the induced changes.
In trees, electrically induced action potentials were measured in willow (Salix viminalis) shoots, showing that calcium influx as well as potassium and chloride efflux are involved in the propagation of signals within the phloem (Fromm and Spanswick, 1993). Using a vibrating electrode in combination with the standard microelectrode technique, an apparent efflux of anions and cations of 200 to 700 pmol cm–2 per action potential was assessed in willow roots (Fromm et al., 1997). These changes in the cellular ion concentrations may be important in intracellular signaling, whereas communication over long distances in trees may be achieved through phloem-transmitted electrical signals. Bridging long distances, these rapid signals possess the capacity for coordinating physiological activities in trees. To ascertain potential functions of long-distance electrical signals in trees that may have a bearing on photosynthesis, noninvasive techniques were employed in this study on poplar (Populus trichocarpa, Populus tremula x P. tremuloides) trees about 80 cm tall. After flame wounding, electrical signals were measured in the phloem by means of the aphid technique (Fromm and Eschrich, 1989). Care was taken to avoid any kind of puncturing stress in the trees, which were grown under nonlimiting nutrient supply or under conditions of calcium deficiency. In parallel, leaf responses were measured in gas exchange (via porometry) and chlorophyll fluorescence (by a two-dimensional imaging approach; Koziolek et al., 2004) at various distances from the wounding site. The hypothesis was tested that (1) heat-induced long-distance electrical signaling affects photosynthesis, and (2) calcium deficiency reduces the capacity for signal transmission. Since intracellular calcium is one of the major elements in the signal transduction pathways of plant cells (Okihara et al., 1991; Trewavas, 2000), calcium-deficient trees were expected to exhibit a difference in behavior.

RESULTS



Long-Distance Electrical Signaling
To detect electrical signals in poplar shoots, the membrane potential was measured either in the leaf mesophyll or in the phloem via severed aphid stylets, at a point in the upper stem (Fig. 1, electrode B) or at the first mature leaf (electrode A). The resting potential of the measured phloem cells ranged between –116 mV to –165 mV in 10 experiments with different plants and was similar to the sieve tube potentials in M. pudica (Fromm, 1991) and maize (Zea mays; Fromm and Bauer, 1994). In agreement with the common observation that electric activity in plants can be incited through contact with ice water, such stimulation of the tip of leaf 1 resulted, in the sieve tubes, in a basipetally propagating electrical signal with a hyperpolarizing amplitude of 25 mV (recorded at either microelectrode position; Fig. 2, top graph). Transmission velocity was 4 to 8 mm s–1. By contrast, when stimulating the lower part of the stem with ice water at the position of leaf 4, action potentials with depolarizing amplitudes of 12 to 20 mV and velocities similar to those reported above were recorded at the two microelectrode positions (Fig. 2, bottom graph). The original resting potential was reestablished about 3 to 4 min after depolarization, thus indicating typical criteria inherent to action potentials.


View larger version (57K):
In this window
In a new window

Figure 1. Experimental arrangement of electrical potential recordings. A, The plant was heat stimulated for 3 s by the flame of a lighter either at the tip of leaf 1 or at the base of leaf 4. Cold stimulation (chilling) was applied to the tip of leaf 1 or the lower stem. To disrupt the propagation of electrical signals, a coldblock of 4°C was applied to the stem between the second and third mature leaf underneath the apex. Coldblocking is not equivalent to cold stimulation (chilling) where ice water is rapidly applied for a few seconds, whereas the coldblock at the stem remains in situ permanently in order to block the excitability of the plasma membranes. B, top, An aphid sucking on the underside of a leaf, in the phloem of the primary vein. B, bottom, After the aphid had been severed from its mouthparts by a microscope laser, the stylet stump exuded sieve tube sap to which the tip of a microelectrode was attached by using a micromanipulator.




View larger version (19K):
In this window
In a new window

Figure 2. Electrical signals following cold stimulation of detached poplar shoots. As described in Figure 1, the plant was cold stimulated by ice water (4°C) for several seconds at the tip of leaf 1 or at the mid of the stem. Electrical signals were monitored in leaf 1 (electrode A) or in the upper stem (electrode B). The arrows denote the instant of stimulation. Upward and downward deflections signify depolarization and hyperpolarization of the membrane potential respectively. The curves are typical examples from five measurements with different plants.


Flame stimulation of the tip of leaf 1 evoked a propagating electrical signal with a hyperpolarizing amplitude of approximately 25 mV as measured in the phloem by electrodes A and B as well as in the mesophyll of the flame-wounded leaf, after electrode A had been inserted into the leaf mesophyll (Fig. 3, top graph). Signal transmission velocity in basipetal direction was 1 to 2 mm s–1. Following flame stimulation of the base of leaf 4, either electrode recorded an irregularly shaped propagating electrical signal with an amplitude of over +50 mV in the phloem as well as in the mesophyll of leaf 1 (Fig. 3, central graph). In contrast to the action potentials evoked by stem chilling (Fig. 2, bottom graph), the flame-induced signals reflected failure in reestablishing the original resting potential, indicating that the voltage changes are different from action potentials. The transmission velocity was 1 to 2 mm s–1 in acropetal direction and, hence, similar to that in basipetally propagating flame-induced signals. To disrupt the propagation of electrical signals, a coldblock of 4°C was applied to the stem between the second and third mature leaf underneath the apex. In contrast to the unchilled plants, the resting potential of the phloem depolarized only slightly at either electrode (Fig. 3, bottom graph), indicating a strong decrease in signal amplitudes.


View larger version (22K):
In this window
In a new window

Figure 3. Responses of the membrane potential of the phloem and the mesophyll to flame stimulation of leaf 1 and leaf 4. Stimulation of the tip of leaf 1 evoked the propagation of electrical signals of a hyperpolarized nature (top graph), while stimulation of leaf 4 generated signals of a depolarized nature (central graph). Applying a coldblock to the stem between leaves 2 and 3 significantly reduced the amplitude of the transmitted signal (bottom graph). The curves are typical examples from five measurements with different plants.


It is remarkable that, independent of the type of stimulation (chilling, heating), signals moving in basipetal direction change toward the negative direction, which contrasts with the opposite response in acropetally traveling signals. The Ca2+-influx/Cl–-efflux/K+-efflux sequence in signal generation may explain the depolarizing direction, but different fluxes are needed to explain signals of the negative sign, such as K+-efflux preceding Cl–-efflux. The latter was proved by blocking K+ channels with 1 mM tetraethylammoniumchloride (TEA+), acting on both sides of the membranes (Wong and Adler, 1986) and applied to the artificial pond water (APW) medium of the stem (Fig. 1). After 1 d, the inhibitor had been distributed within the plant. When these plants were flame stimulated at the tip of leaf 1, no hyperpolarizing signal was measured in the phloem (Fig. 4, top graph). Obviously, transient hyperpolarization is caused by the transient opening of K+ channels in the plasma membrane.


View larger version (27K):
In this window
In a new window

Figure 4. Typical responses of the membrane potential in phloem cells upon flame stimulation in TEA+-treated as well as calcium-deficient plants. Upon application of TEA+, a K+ channel blocker, flame stimulation of leaf 1 did not evoke an electrical signal (top graph). Also in Ca-deficient plants, stimulation of leaf 1 failed to generate any signal (central graph), and wounding of leaf 4 caused weak changes in the membrane potential (bottom graph).


As the ionic mechanism of excitation is based on the movement of potassium, chloride, and calcium (Tazawa et al., 1987; Fromm and Spanswick, 1993), the objective of experiments involving the stimulation of calcium-deficient plants was to characterize the excitation mechanism in poplar. The Ca-deficient plants were distinctly smaller and their leaf size was reduced compared with that in nonlimited plants. Flame wounding of the tip of leaf 1 evoked no detectable electrical signals in the phloem at either electrode position (Fig. 4, central graph). However, signals with strongly reduced amplitudes were measured in acropetal direction after flame stimulation of leaf 4 (Fig. 4, bottom graph), perhaps reflecting the physiological significance of the upwards transmission of electrical signals in plant life.

Chlorophyll Fluorescence Imaging
No photosynthetic response was found upon stimulation of leaves or stem with ice water. However, flaming of the tip of leaf 1 caused a substantial decrease in the electron quantum yield of PSII (Fig. 5A). This effect occurred 80 s after flame stimulation in the intercostal regions of the central part of the lamina, at a distance of 3 cm from the leaf tip. The inhibitory response spread basipetally throughout the leaf, showing a delay in the arrival of the electrical signal at electrode B (approximately 60 s after flame wounding of the leaf tip) in relation to electrode A (after only 2 s; Fig. 3, top graph). In the flame-stimulated leaf, the fluorescence response reached a minimum after about 300 s, prior to incipient recovery.



View larger version (216K):
In this window
In a new window

Figure 5. Spatiotemporal changes of PSII electron quantum yield, ΔF/F'm assessed by chlorophyll fluorescence imaging. The imaged area (l, 22 mm; w, 17 mm) covers the center of leaf 1. A, The tip of leaf 1 was flame stimulated at a distance of 3 cm. Times are given in relation to the instant of injury (at time 0). Changes in ΔF/F'm took 80 s to become apparent. A false-color shift from blue to yellow in the intervein area (equivalent to a lowering of PS electron quantum yield from 0.6 to about 0.2) indicates the decrease in the electron quantum yield of PSII. B, Leaf 4 was wounded at a distance of over 11 cm from the imaged area. Changes in fluorescence were not visible before 240 s had elapsed. After about 420 and 600 s, the left and right side of the leaf, respectively, reached a minimum in the fluorescence response (white and black arrow, respectively). The bar translates the false color code into values of ΔF/F'm.


Flaming of leaf 4 caused a decrease in the electron quantum yield of PSII in leaf 1, evidence of which was found in the leaf veins 240 s of the latter after flame stimulation, and indicating that the signal spreads via the veins into the mesophyll (arrival after 300 s; Fig. 5B). The inhibitory response propagated acropetally throughout the leaf and was delayed in relation to the electrical signal that arrived in leaf 1 only 80 s after flaming of leaf 4 (Fig. 3, central graph). A minimum in the time course of the fluorescence response was reached in the left half of the lamina of leaf 1 after about 420 s (Fig. 5B, white arrow) prior to incipient recovery. However, it took about 600 s in the right half of the leaf for the minimum to occur (black arrow).
The highly resolved time course of the fluorescence response shows the decrease in PSII quantum yield to occur simultaneously in both the veins and the intercostal regions of leaf 1 taking 160 s from the time of flame stimulation of the leaf tip (distance approximately 5 cm; Fig. 6A). By contrast, a distinct two-step response was observed in leaf 1 during acropetal signaling when leaf 4 was flame stimulated, as the response in veins was faster and stronger than in the intercostal regions (Fig. 6B). In these latter areas, the transient response was delayed by about 60 s compared with the occurrence in the veins. The differences in the response between long-distance signaling along the shoot (Fig. 6B) and short-distance signaling within the same leaf (Fig. 6A) reveal, in the first case, that the signals spread into the mesophyll via the veins, whereas in the latter case, signals travel across veins and mesophyll at similar velocities.


View larger version (52K):
In this window
In a new window

Figure 6. Kinetic changes of PSII electron quantum yield, ΔF/F'm, measured by fluorescence imaging at representative sites within leaf 1 after flame wounding of A, its leaf tip, and B, the base of leaf 4 at time 0. Blue and red symbols represent vein and intervein field areas, respectively. Data shown are means ± SD, n = 5. Parts of imaged areas showing representative vein and intervein field areas used for kinetic analysis are shown below the graphs.



Leaf Gas Exchange
In leaf 1, the net CO2 uptake rate (JCO2) sharply dropped to about compensation 30 s after flame stimulation of the leaf tip and stayed there for about 90 s before subsequent recovery (Fig. 7A). At the same time, the stomatal conductance (gH2O) remained stable, indicating the absence of stomatal movements. The finding of a photosynthetic reaction not showing any reaction in the gH2O is astonishing. However, the internal CO2 concentration increased transiently (data not shown), indicating that the relationship between JCO2 and gH2O is correct. A similar response occurred in leaf 1, when leaf 4 was flame stimulated at a distance of over 10 cm. At 120 s after stimulation, the JCO2 decreased immediately to compensation, staying there for 100 s and then recovering almost completely (Fig. 7B). Chilling the stem slowly down to a temperature of +4°C (coldblock) between leaves 2 and 3 showed that the gas exchange of leaf 1 remained unaltered (Fig. 7C), concomitant with the suppression of the long-distance electrical signal (Fig. 3, bottom graph). This coincidence indicates that the electrical signal has a modifying impact on CO2 uptake. Experimentation with calcium-deficient plants did not result in any significant response in gas exchange after flame stimulation of leaf 4 (Fig. 7D), which is consistent with the distinctly reduced amplitude of the recorded electrical signal (Fig. 4, bottom graph).


View larger version (21K):
In this window
In a new window

Figure 7. Typical responses of JCO2 and gH2O of leaf 1 upon heat stimulation of its tip (A) or of leaf 4 (B). The arrows denote the instant of injury. For heat stimulation of the leaf tip, mean of JCO2 before stimulation was 1.35 ± 0.15 and mean of JCO2 at transient minimum was –0.2 ± 0.2, while mean of gH2O was 105 ± 15, n = 5. For heat stimulation of leaf 4, mean of JCO2 before stimulation was 1.38 ± 0.21 and mean of JCO2 at transient minimum was 0.48 ± 0.5, while mean of gH2O was 108 ± 17, n = 5. C, Chilling the stem with ice water evoked no response in gas exchange. D, No significant response in gas exchange was recorded in calcium-deficient plants following flame stimulation of leaf 4.


To determine the velocity of chemical signaling, the petioles of excised leaves were fed with 14C-labeled Suc. After a 600-s translocation, macroautoradiography showed that the primary vein became labeled and 14C-Suc extended from the vein across the entire lamina (data not shown). Since the velocity of electrical signal transmission following flame stimulation was 1 to 2 mm s–1 in both acropetal and basipetal direction, autoradiography proved chemical signaling to be much too slow to account for the photosynthetic response after flame stimulation
DISCUSSION


The results presented here show that electrical signals propagated over long distances as well as short distances are capable of modifying photosynthesis in trees. Previous studies had reported on the capacity of many plant species to generate and transmit action potentials as well as variation potentials (Pickard, 1973; Davies, 1987). In ecophysiology, electrical signals play a key role in the communication of environmental stimuli within plants, apparently acting within and across cells, tissues, and organs (Volkov and Mwesigwa, 2001). For instance, the pesticide 2,4-dinitrophenol, considered an environmental problem in agriculture as well as a human health hazard, induced fast action potentials and decreased the variation potential in soybean (Glycine max; Mwesigwa et al., 2000). Pentachlorophenol too, a known pollutant and uncoupler of oxidative phosphorylation, induced action potentials in soybean (Volkov et al., 2000). Recent experiments on signal transmission in M. pudica demonstrated that a link exists between flame-induced electrical signals and photosynthetic responses, as inferred from two-dimensional imaging analysis of chlorophyll fluorescence in combination with gas exchange assessment (Koziolek et al., 2004).
This study on poplar confirms the latter findings for trees. Moreover, our results demonstrate that different stimulation types and positions incite characteristic electrical signals, each with a specific influence on photosynthesis. We used the aphid technique to measure electrical signals in phloem cells that share fundamental properties with nerve cells in animals, i.e. the existence of excitable membranes by which electrical excitations can be transmitted from cell to cell. In poplar, the nature of the signal depends on the traveling direction. Basipetally propagating signals (induced by chilling as well as flaming) showed negative voltage changes, whereas acropetally moving signals were characterized by transient membrane depolarization (Figs. 2 and 3). The ionic mechanism of action potentials in plants, i.e. depolarizing signals with positive sign, is based on calcium influx as well as initial chloride efflux followed by potassium efflux (Beilby and Coster, 1979; Kikuyama and Tazawa, 1983; Lunevsky et al., 1983; Tsutsui et al., 1986; Kikuyama, 1987; Fromm and Spanswick, 1993). To explain signals of the negative sign, we blocked K+ channels by TEA+ and found that no basipetally transmitted signals were induced after flame stimulating the tip of leaf 1 (Fig. 4, top graph), indicating that K+ efflux causes the hyperpolarization of basipetally transmitted signals. A similar mechanism is reported for the unicellular green alga Eremosphaera viridis, where a sudden blockage of photosynthetic electron transport by darkening causes a transient hyperpolarization of the plasma membrane (Schönknecht et al., 1998). In Eremosphaera, the transient hyperpolarization is due to the opening of K+ channels that is caused by a rapid transient elevation of the cytosolic free calcium concentration. In poplar plants grown under calcium deficiency, phloem cell excitability was completely inhibited after flaming of the leaf tip (compare with Fig. 4). This result indicates first, that calcium channels are involved in the induction of basipetally transmitted electrical signals, confirming the observation that calcium channels and transporters form the basis of a complex calcium signaling network (Trewavas and Malho, 1997; Trewavas, 2000). Second, in optimum nutrient-supplied plants, blockage of the signal by TEA+ indicates that K+ efflux is responsible for the occurrence of hyperpolarization, most likely induced by calcium as reported for Eremosphaera (Schönknecht et al., 1998).
In the poplars of this study, electrical signals were induced by chilling as well as by flame wounding. As regards chilling, amplitudes and dynamics of the acropetally transmitted signals were typical of plant action potentials, as previously demonstrated in maize (Fromm and Fei, 1998) and, for trees, in willow (Fromm and Spanswick, 1993). Action potentials are generally self-amplifying signals propagated at constant velocity. It is assumed that they are dependent on voltage-gated ion channels and are capable of propagating through any living cells sharing common membranes, but they are most evident in sieve tubes. In contrast to action potentials, flame stimulation causes the appearance of a slowly moving, irregular, so-called variation potential (Sibaoka, 1966; Sibaoka, 1969). There is widespread agreement in the literature that the variation potential is not a self-propagating signal but, rather, a local electrical response to the underlying passage of chemical substances released from the wound site and transmitted through the xylem by hydraulic dispersal. Malone (1992) reported on hydraulic signals in wheat leaves, traveling at a velocity of at least 100 mm s–1 from the site of stimulation. However, using a combination of electrometer and laser-interferometer, Tinz-Füchtmeier and Gradmann (1990) were unable to confirm the hydraulic conductance of excitation in Mimosa.
In poplar, hydraulic signals may play a role in the generation of electrical excitation. Since we managed to measure flame-induced signals in the phloem, the question arises whether pressure or chemical changes in the xylem can activate ion channels in the phloem, making it appear as if a hydraulically induced variation potential were passing through the phloem. However, as the electric signal transmission was disrupted after applying a coldblock to the stem and gas exchange did not respond to flame wounding, it is thought unlikely that hydraulic events play any role in long-distance signaling in poplar. Moreover, variation potentials certainly depend on the prevailing water status of the plant. When shoot water status is saturated, as in the case of the well-watered plants used in this study, xylem tension becomes negligible and variation potentials should not travel at all. We are therefore convinced that the phloem-transmitted flame-induced signals in poplar are self-propagating signals, independent of chemicals traveling through the xylem.
Concerning chemical signaling in the phloem that spreads from the stimulation site through the sieve tubes, the thought cannot be dismissed that photosynthesis is affected by chemical signals. Canny (1975) reported transport velocities in the phloem to range between 50 and 100 cm h–1, i.e. too slow to play a key role in the studied poplar trees. The first response in gas exchange upon wounding of the leaf tip took 30 s to travel a distance of 4 cm, whereas a chemical signal would have moved across a distance of no more than 0.4 to 0.8 cm in 30 s. In consistency with these findings, autoradiography confirmed that chemical signaling is too slow to account for the photosynthetic response (data not shown).
With regard to the photosynthetic response, only flame-evoked signals caused photosynthetic changes, whereas chill-induced signals had no impact on photosynthesis. Interestingly, the noninvasive imaging analysis of chlorophyll fluorescence revealed that short-distance signals within leaves after flame stimulation at the tip cause a simultaneous decrease in electron quantum yield of PSII in both veins and intercostal regions (compare with Fig. 6A). By contrast, long-distance signals, again generated by flame wounding of leaves but arriving after having traveled across the plant, reduce the electron quantum yield of PSII in the veins first and in the intercostal regions afterward (Fig. 6B), hence suggesting that the signal spreads via the veins into the mesophyll. Reduction in photosynthesis upon impact by electrical signals is also known for M. pudica (Koziolek et al., 2004) and tomato plants (Herde et al., 1999). Since the PSII quantum yield decreased as well as the JCO2, it can be excluded that the latter is only based on increased respiration. However, the mechanism underlying photosynthetic limitation upon impact by electrical signals requires further clarification. Subcellular alterations in ion fluxes may be involved in the photosynthetic response so that attention should focus on the translocation of ions in mesophyll cells and their chloroplasts. It is reported that enzymes in the cell wall, the plasmalemma, and the cytoplasm can show modified activities during local changes in ion concentrations (Davies, 1987). Bulychev et al. (1986, 1987) reported changes in the chlorophyll fluorescence of intact chloroplasts induced by shifts in membrane potential. The rapid impact on chlorophyll fluorescence and CO2 uptake in poplar is likely, therefore, to be caused by local changes in ion concentrations. For spinach, evidence is presented for a direct involvement of calcium in O2 formation of PSII (Vrettos et al., 2001), indicating a possible role for calcium in the photosynthetic response. In summary, this study provides evidence that rapidly evoked and phloem-transmitted electrical signals can affect photosynthesis over long distances in trees. However, the intracellular controlling points in the leaf mesophyll will be the subject of further investigations.

MATERIALS AND METHODS



Plant Material and Growth Conditions
One-year-old plants of Populus trichocarpa cv Trichobel were grown in the summer from cuttings under standard greenhouse conditions, at 25°C and 85% relative humidity, in sandy culture medium. In addition, 8-week-old plants of Populus tremula x P. tremuloides Michx. were grown hydroponically but under otherwise identical greenhouse conditions. The latter plants were provided with macro- and micronutrients in a modified Hoagland solution (Hoagland and Arnon, 1950) containing either low Ca2+ (0.1 mM) or full-strength Ca2+ (5 mM) concentration. Experiments were performed using fully developed shoots from both groups of plants. Effects of calcium deficiency on electrical signaling were studied using the hydroponically grown plants (P. tremula x P. tremuloides Michx.); all other experiments were performed on P. trichocarpa cv Trichobel.

Electrical Potential Measurements
When about six leaves had developed, plants were cut from their roots and transferred to a Faraday cage for experimentation. They had to be cut from their roots in order to detect the membrane potential by measurements from two sides of the membrane. Microelectrode tips were either inserted into the leaf mesophyll or brought into contact with the exudate droplet on a severed aphid stylet. In the latter case, contact was made at two different positions, either at some point in the upper stem (electrode B) or at the lower side of the first mature leaf underneath the apex (electrode A, Fig. 1). The reference electrode (Ag/AgCl) was immersed into the APW where the cut cross section of the excised stem had also been submerged. The APW was composed of 1.0 mM NaCl, 0.1 mM KCl, 0.1 mM CaCl2, and 1.0 mM MES, adjusted with Tris to a pH value of 6.0.
The electrical potential of the phloem was measured via severed aphid stylets. This involved introducing aphids to a leaf or the stem and allowing them to settle overnight. On the following day, they were severed from their stylets by shots from a laser beam generator (Beck, Neu-Isenburg, Germany), connected to a Zeiss microscope. Electric potential changes were measured through glass microelectrodes with tip diameters of less than 1 µm, back-filled with 3 M KCl. The microelectrode was clamped in an Ag/AgCl pellet holder (WPI) and connected to a microelectrode preamplifier (input impedance >1012 ohms) to which a WPI amplifier (model 750, WPI, Sarasota, FL) was attached. The response time of the microelectrodes is about 1 s, which is fast enough to measure electrical responses induced by heat and cold stimulation. The electrodes were inserted into the leaf mesophyll or attached to the stylet stump by micromanipulators: electrode A in leaf 1 at a mean distance of 30 to 40 mm to the stimulated tip; electrode B at some point in the stem (Fig. 1). The resistance for an electric current inside the aphid stylet is relatively low (around 109 Ω according to Wright and Fisher, 1981) compared to the high input impedance of the electric equipment used. Recordings were made between the microelectrodes and a reference electrode (in APW and connected to the cut end of the shoot). Measurement results were logged in parallel by chart recorder and computer.

Chlorophyll Fluorescence Imaging
The two-dimensional imaging approach described by Koziolek et al. (2004; Imaging-PAM Chlorophyll Fluorometer, Heinz Walz GmbH, Effeltrich, Germany) was employed to assess the spatiotemporal variations of the quantum yield of energy conversion in PSII (Siebke and Weis, 1995; Rascher et al., 2001). This method allows noninvasive determination of PSII quantum yield by the saturation pulse method (Schreiber et al., 1986; Genty et al., 1989). Blue light (470 nm) is applied to act as pulse-modulated measuring light, actinic illumination, and saturation pulses. The imaged area of leaf 1 was adapted for at least 10 min to a PPFD of 100 µmol m–2 s–1 prior to flame wounding of the leaf tip (same leaf) or an adjacent leaf. Heat treatment involved a flame stimulus lasting 3 s. Saturation pulses were given every 10 s to determine the images of fluorescence yield, F, maximal fluorescence yield, Fm', and PSII quantum yield, ΔF/Fm' = (Fm'– F)/Fm' (for nomenclature, see van Kooten and Snel, 1990).

Leaf Gas Exchange
The gas exchange in the attached leaves was measured using a steady-state CO2/water diffusion porometer (CQP130, Walz, Effeltrich, Germany) at ambient CO2 concentration of about 360 µL L–1 and a relative humidity of approx. 60%, leaf temperature of about 27°C, and PPFD of about 100 µmol m–2 s–1. Leaf 1 was attached to the porometer and its leaf tip or leaf 4 (compare with Fig. 1) were stimulated by flaming. Leaf net-CO2 exchange was calculated based on single leaf area and expressed as JCO2.
 

cannawizard

Well-Known Member
http://www.irnase.csic.es/users/jefer/Articulos en papel JE/AGUA EN SUELO/bingham2001.pdf

........................................................








Root Exudation and Rhizosphere Biology

+ Author Affiliations​
  • 1 Department of Horticulture and Landscape Architecture (T.S.W., H.P.B., J.M.V.),​
  • 2 Cellular and Molecular Biology Graduate Program (J.M.V.), and​
  • 3 Graduate Degree Program in Ecology (J.M.V.), Colorado State University, Fort Collins, Colorado 80523; and​
  • 4 Department of Plant Biology and Plant Biotechnology Center, The Ohio State University, Columbus, Ohio 43210 (E.G.)​
Our understanding of the biology, biochemistry, and genetic development of roots has considerably improved during the last decade (Smith and Fedoroff, 1995; Flores et al., 1999;Benfey and Scheres, 2000). In contrast, the processes mediated by roots in the rhizosphere such as the secretion of root border cells and root exudates are not yet well understood (Hawes et al., 2000). In addition to the classical roles of providing mechanical support and allowing water/nutrient uptake, roots also perform certain specialized roles, including the ability to synthesize, accumulate, and secrete a diverse array of compounds (Flores et al., 1999). Given the complexity and biodiversity of the underground world, roots are clearly not passive targets for soil organisms. Rather, the compounds secreted by plant roots serve important roles as chemical attractants and repellants in the rhizosphere, the narrow zone of soil immediately surrounding the root system (Estabrook and Yoder, 1998; Bais et al., 2001). The chemicals secreted into the soil by roots are broadly referred to as root exudates. Through the exudation of a wide variety of compounds, roots may regulate the soil microbial community in their immediate vicinity, cope with herbivores, encourage beneficial symbioses, change the chemical and physical properties of the soil, and inhibit the growth of competing plant species (Nardi et al., 2000; Fig. 1A). The ability to secrete a vast array of compounds into the rhizosphere is one of the most remarkable metabolic features of plant roots, with nearly 5% to 21% of all photosynthetically fixed carbon being transferred to the rhizosphere through root exudates (Marschner, 1995).


View attachment 1789388


Fig. 1.A, Representation of the complex interactions mediated by root exudates that take place in the rhizosphere between plant roots and other organisms. Organisms are not drawn to scale. QS, quorum sensing. B, In vitro culture of oca (Oxalis tuberosa) grown in sterile liquid medium under UV light exposure. C, Chemical structure of harmine as determined by1H and C13 NMR analysis. D, Fluorescent root exudates from O. tuberosa were observed bound to the blue germination paper under UV light exposure. E, Soil samples showing fluorescence obtained from greenhouse-grown oca plants. Samples were taken 5 cm from the stem girth of the plant, and the numbers (1–8) denote the depth by every 1 cm toward the top-layer soil. In vitro-grown oca plants and soil samples collected from oca's rhizosphere were visualized for blue-purplish fluorescence under UV light exposure with a short wave of UV approximately 254 nm.







Although root exudation clearly represents a significant carbon cost to the plant, the mechanisms and regulatory processes controlling root secretion are just now beginning to be examined. Root exudates have traditionally been grouped into low- and high-Mr compounds. However, a systematic study to determine the complexity and chemical composition of root exudates from diverse plant species has not been undertaken. Low-Mr compounds such as amino acids, organic acids, sugars, phenolics, and various other secondary metabolites are believed to comprise the majority of root exudates, whereas high-Mr exudates primarily include mucilage (high-Mr polysaccharides) and proteins.

The rhizosphere is a densely populated area in which the roots must compete with the invading root systems of neighboring plant species for space, water, and mineral nutrients, and with soil-borne microorganisms, including bacteria, fungi, and insects feeding on an abundant source of organic material (Ryan and Delhaize, 2001). Thus, root-root, root-microbe, and root-insect communications are likely continuous occurrences in this biologically active soil zone, but due to the underground nature of roots, these intriguing interactions have largely been overlooked. Root-root and root-microbe communication can either be positive (symbiotic) to the plant, such as the association of epiphytes, mycorrhizal fungi, and nitrogen-fixing bacteria with roots; or negative to the plant, including interactions with parasitic plants, pathogenic bacteria, fungi, and insects. Thus, if plant roots are in constant communication with symbiotic and pathogenic organisms, how do roots effectively carry out this communication process within the rhizosphere?

A large body of knowledge suggests that root exudates may act as messengers that communicate and initiate biological and physical interactions between roots and soil organisms. This update will focus on recent advancements in root exudation and rhizosphere biology.



Next Section

ROOT-RHIZOSPHERE COMMUNICATION



Survival of any plant species in a particular rhizosphere environment depends primarily on the ability of the plant to perceive changes in the local environment that require an adaptive response. Local changes within the rhizosphere can include the growth and development of neighboring plant species and microorganisms. Upon encountering a challenge, roots typically respond by secreting certain small molecules and proteins (Stintzi and Browse, 2000;Stotz et al., 2000). Root secretions may play symbiotic or defensive roles as a plant ultimately engages in positive or negative communication, depending on the other elements of its rhizosphere. In contrast to the extensive progress in studying plant-plant, plant-microbe, and plant-insect interactions that occur in aboveground plant organs such as leaves and stems, very little research has focused on root-root, root-microbe, and root-insect interactions in the rhizosphere. The following sections will examine the communication process between plant roots and other organisms in the rhizosphere.

Root-Root Communication



In natural settings, roots are in continual communication with surrounding root systems of neighboring plant species and quickly recognize and prevent the presence of invading roots through chemical messengers. Allelopathy is mediated by the release of certain secondary metabolites by plant roots and plays an important role in the establishment and maintenance of terrestrial plant communities. It also has important implications for agriculture; the effects may be beneficial, as in the case of natural weed control, or detrimental, when allelochemicals produced by weeds affect the growth of crop plants (Callaway and Aschehoug, 2000). A secondary metabolite secreted by the roots of knapweed (Centaurea maculosa) provides a classic example of root exudates exhibiting negative root-root communication in the rhizosphere. Recently, Bais et al. (2002c) identified (±)-catechin as the root-secreted phytotoxin responsible for the invasive behavior of knapweed in the rhizosphere. Interestingly, (−)-catechin was shown to account for the allelochemical activity, whereas (+)-catechin was inhibitory to soil-borne bacteria (Bais et al., 2002c ). In addition to racemic catechin being detected in the exudates of in vitro-grown plants, the compound was also detected in soil extracts from knapweed-invaded fields, which strongly supported the idea that knapweed's invasive behavior is due to the exudation of (−)-catechin. Moreover, this study established the biological significance of the exudation of a racemic compound such as catechin, demonstrating that one enantiomer can be responsible for the invasive nature of the plant, whereas the other enantiomer can contribute to plant defense.

Although studies have reported the biosynthesis of the common enantiomer (+)-catechin, little is known regarding the synthesis of (−)-catechin or (±)-catechin as natural products. One possibility is that (+)-catechin production is followed by racemization in the root or during the exudation process. Alternatively, there could be a deviation from the normally observed stereo- and enantiospecific biosynthesis steps. The flavonols kaempferol and quercetin are generally perceived as final products, rather than intermediates, in the pathway (Winkel-Shirley, 2001). The correlation of these experiments to the root exudation process has yet to be determined, but the data should provide a starting point for further studies on the characterization of specific committed steps in the synthesis of racemic catechin in knapweed roots.

The above example demonstrates how plants use root-secreted secondary metabolites to regulate the rhizosphere to the detriment of neighboring plants. However, parasitic plants often use secondary metabolites secreted from roots as chemical messengers to initiate the development of invasive organs (haustoria) required for heterotrophic growth (Keyes et al., 2000). Some of the most devastating parasitic plants of important food crops such as maize (Zea mays), sorghum (Sorghum bicolor), millet (Panicum milaceum), rice (Oryza sativa), and legumes belong to the Scrophulariaceae, which typically invade the roots of surrounding plants to deprive them of water, minerals, and essential nutrients (Yoder, 2001). It has been reported that certain allelochemicals such as flavonoids,p-hydroxy acids, quinones, and cytokinins secreted by host roots induce haustorium formation (Estabrook and Yoder, 1998; Yoder, 2001), but the exact structural requirements of the secreted compounds for haustorium induction is not fully understood.



Root-Microbe Communication



Root-microbe communication is another important process that characterizes the underground zone. Some compounds identified in root exudates that have been shown to play an important role in root-microbe interactions include flavonoids present in the root exudates of legumes that activate Rhizobium meliloti genes responsible for the nodulation process (Peters et al., 1986). Although the studies are not yet conclusive, these compounds may also be responsible for vesicular-arbuscular mycorrhiza colonization (Becard et al., 1992, 1995; Trieu et al., 1997). In contrast, survival of the delicate and physically unprotected root cells under continual attack by pathogenic microorganisms depends on a continuous “underground chemical warfare” mediated by secretion of phytoalexins, defense proteins, and other as yet unknown chemicals (Flores et al., 1999).

The unexplored chemodiversity of root exudates is an obvious place to search for novel biologically active compounds, including antimicrobials. For instance, Bais et al. (2002b)recently identified rosmarinic acid (RA) in the root exudates of hairy root cultures of sweet basil (Ocimum basilicum) elicited by fungal cell wall extracts from Phytophthora cinnamoni. Basil roots were also induced to exude RA by fungal in situ challenge withPythium ultimum, and RA demonstrated potent antimicrobial activity against an array of soil-borne microorganisms includingPseudomonas aeruginosa (Bais et al., 2002b). Similar studies by Brigham et al. (1999) withLithospermum erythrorhizon hairy roots reported cell-specific production of pigmented naphthoquinones upon elicitation, and other biological activity against soil-borne bacteria and fungi. Given the observed antimicrobial activity of RA and naphthoquinones, these findings strongly suggest the importance of root exudates in defending the rhizosphere against pathogenic microorganisms. Moreover, the aforementioned studies complement earlier research that mainly focused on the regulation and production of these compounds by providing valuable insights into the biological importance of RA and shikonin.

Both Gram-negative and -positive bacteria, including important plant pathogenic bacteria such as Erwinia spp.,Pseudomonas spp., and Agrobacterium spp., possess quorum-sensing systems that control the expression of several genes required for pathogenicity (for review, see Fray, 2002). Quorum sensing is a form of cell-cell communication between bacteria mediated by small diffusible signaling molecules (autoinducers); these are generally acylated homo-Ser lactones (AHLs) for Gram-negative bacteria and peptide-signaling molecules for Gram-positive bacteria. Upon reaching a threshold concentration at high-population densities, an auto-inducer then activates transcriptional activator proteins that induce specific genes. Thus, intercellular signals enable a bacterial population to control the expression of genes in response to cell density. A recent review by Fray (2002) reported that AHL-producing transgenic tobacco plants restored pathogenicity to an avirulent AHL-deficient Erwinia carotovora mutant. Root exudates from pea (Pisum sativum) seedlings were found to contain several bioactive components that mimicked AHL signals in well-characterized bacterial reporter strains, stimulating AHL-regulated behaviors in some strains while inhibiting such behaviors in others. The chemical nature of such active mimic secondary metabolites is currently unknown (Teplitski et al., 2000; Knee et al., 2001). However, it was also reported that crude aqueous extracts from several plant species exhibited AHL inhibitory activity. Thus, it is possible that roots may have developed defense strategies by secreting compounds into the rhizosphere that interfere with bacterial quorum-sensing responses such as signal mimics, signal blockers, and/or signal-degrading enzymes, but future studies are required to isolate and characterize these compounds.



Root-Insect Communication



The study of plant-insect interactions mediated by chemical signals has largely been confined to leaves and stems, whereas the study of root-insect communication has remained largely unexplored due to the complexity of the rhizosphere and a lack of suitable experimental systems. However, root herbivory by pests such as aphids can cause significant decreases in yield and quality of important crops including sugar beet (Beta vulgaris), potato (Solanum tuberosum), and legumes (Hutchison and Campbell, 1994). One attempt to study root-insect communication was developed by Wu et al. (1999) using an in vitro coculture system with hairy roots and aphids. In this study, it was observed that aphid herbivory reduced vegetative growth and increased the production of polyacetylenes, which have been reported to be part of the phytoalexin response (Flores et al., 1988). In a more recent study, Bais et al. (2002a) reported the characterization of fluorescent β-carboline alkaloids from the root exudates of O. tuberosa (oca). The main fluorescent compounds were identified as harmine (7-methoxy-1-methyl-β-carboline) and harmaline (3, 4-dihydroharmine; Bais et al., 2002a; Fig. 1, B–E). In addition to their fluorescent nature, these alkaloids exhibit strong phototoxicity against a polyphagous feeder,Trichoplusia ni, suggesting their insecticidal activity may be linked to photoactivation (Larson et al., 1988). The Andean highlands, where O. tuberosa is primarily cultivated, are subjected to a high incidence of UV radiation, and it was observed that the strongest fluorescence intensity occurred with oca varieties that showed resistance to the larvae ofMycrotrypes spp., the Andean tuber weevil (Flores et al., 1999). These data suggest that UV light penetrating soil layers could photoactivate fluorescent β-carboline alkaloids secreted by oca roots to create an insecticidal defense response.





Previous SectionNext Section

ALTERATION OF SOIL CHARACTERISTICS THROUGH EXUDATION



As a consequence of normal growth and development, a large range of organic and inorganic substances are secreted by roots into the soil, which inevitably leads to changes in its biochemical and physical properties (Rougier, 1981). Various functions have been attributed to root cap exudation including the maintenance of root-soil contact, lubrication of the root tip, protection of roots from desiccation, stabilization of soil micro-aggregates, and selective adsorption and storage of ions (Griffin et al., 1976;Rougier, 1981; Bengough and McKenzie, 1997; Hawes et al., 2000). Root mucilage is a reasonably studied root exudate that is believed to alter the surrounding soil as it is secreted from continuously growing root cap cells (Vermeer and McCully, 1982; Ray et al., 1988; McCully, 1995; Sims et al., 2000). Soil at field capacity typically possesses a matric potential of −5 to −10 kPa (Chaboud and Rougier, 1984). It has been speculated that as the soil dries and its hydraulic potential decreases, exudates will subsequently begin to lose water to soil. When this occurs, the surface tension of the exudates decreases and its viscosity increases. As the surface tension decreases, the ability of the exudates to wet the surrounding soil particles will become greater. In addition, as viscosity increases, the resistance to movement of soil particles in contact with exudates will increase, and a degree of stabilization within the rhizosphere will be achieved. For instance, McCully and Boyer (1997)reported that mucilage from the aerial nodal roots of maize has a water potential of −11 Mpa, indicating a large capacity for water storage when fully hydrated, whereas the mucilage loses water to the soil as it begins to dry.

This speculation supports the idea that root exudates could play a major role in the maintenance of root-soil contact, which is especially important to the plant under drought and drying conditions, when hydraulic continuity will be lost. The largest, most coherent soil rhizosheaths are formed on the roots of grasses in dry soil (Watt et al., 1994). However, sheath formation requires fully hydrated exudates to permeate the surrounding soil particles that are then bonded to the root and each other as the mucilage dries. Young (1995) found that rhizosheath soil was significantly wetter than bulk soil and suggested that exudates within the rhizosheath increase the water-holding capacity of the soil. Furthermore, it has recently been proposed that in dry soil, the source of water to hydrate and expand exudates is the root itself. Modern cryo-scanning microscopy has helped researchers determine that the rhizosheath of a plant is more hydrated in the early morning hours compared with the midday samplings (McCully and Boyer, 1997). This implies that the exudates released from the roots at night allow the expansion of the roots into the surrounding soil. When transpiration resumes, the exudates begin to dry and adhere to the adjacent soil particles. Thus, the rhizosheath is a dynamic region, with cyclic fluctuations in hydration content controlled to some extent by roots.

Taken together, these studies indicate that root exudation plays a major role in maintaining root-soil contact in the rhizosphere by modifying the biochemical and physical properties of the rhizosphere and contributing to root growth and plant survival. However, the exact fate of exuded compounds in the rhizosphere, and the nature of their reactions in the soil, remains poorly understood.



Previous SectionNext Section

CELLULAR MECHANISMS OF ROOT EXUDATION



Subcellular Trafficking of Exuded Metabolites



Despite the ecophysiological significance of plant-secreted compounds and the large number of compounds that plant cells produce, very little is currently known about the molecular mechanisms for the trafficking of phytochemicals. In at least some plants, channels are likely to be involved in the secretion of organic acids normally present at high levels in the cytoplasm. A good example is provided by the exudation of citrate, malate, and related organic acids by maize and wheat (Triticum aestivum) in response to high Al3+ concentrations (Ma et al., 2001). However, plants have the potential to express 100,000 compounds, primarily derived from secondary metabolism (Verpoorte, 2000), many of them with cytotoxic activities that would prevent their accumulation in the cytoplasm. The speculation that phytochemicals are transported from the site of synthesis to the site of storage by vesicles or specialized organelles is gaining momentum as evidence accumulates regarding the presence of intracellular bodies in plant cells induced to accumulate large quantities of secondary metabolites (Grotewold, 2001). For example, it has long been known that specific steps of the isoquinoline alkaloid biosynthetic pathway are sequestered in alkaloid vesicles and that pathway intermediates must traffic from one subcellular compartment to another by mechanisms that prevent their free diffusion in the cytosol (Facchini, 2001). Subcellular inclusions that accumulate 3-deoxy anthocyanidin flavonoid phytoalexins are observed in sorghum leaves infected by the fungusColletotrichum graminicola (Snyder and Nicholson, 1990). These inclusions are similar to the anthocyanoplasts observed in maize cells expressing the C1 and R regulators of anthocyanin accumulation (Grotewold et al., 1998).

Root exudates often include phenylpropanoids and flavonoids, presumably synthesized on the cytoplasmic surface of the endoplasmic reticulum (ER; Winkel-Shirley, 2001). For example, the flavone luteolin, secreted by alfalfa (Medicago sativa) seedlings and seed coats, provides one of the signals that induces the nodulation genes in R. meliloti (Peters et al., 1986). Cytotoxic and antimicrobial catechin flavonoids are secreted by the roots of knapweed plants (Bais et al., 2002c). Although the mechanisms by which these compounds are transported from the ER to the plasma membrane are not known, it is possible that they are transported by ER-originating vesicles that fuse to the cell membrane and release their contents.

Vesicles with the above-described properties and containing green autofluorescent compounds have been identified in maize cells ectopically expressing the P regulator of 3-deoxy flavonoid biosynthesis (Grotewold et al., 1998). These vesicles are likely to originate from the ER, as suggested by the presence of green fluorescence inside specific regions of the ER after treatment with brefeldin A. The vesicles fuse and form large green fluorescent bodies that migrate to the surface of the cell and fuse to the cell membrane and release the green fluorescent compound to the cell wall (Grotewold et al., 1998). Interestingly, the accumulation of the green fluorescence in the cell wall is increased by treatment with Golgi-disrupting agents, such as brefeldin A or monensin, suggesting a trans-Golgi network-independent pathway for the secretion of these compounds. Cultured cells of maize ectopically expressing P also accumulate increased quantities of yellow autofluorescent compounds that are targeted to the central vacuole by subcellular structures that resemble anthocyanoplasts (Grotewold et al., 1998). The use of these autofluorescent compounds, or the fluorescent β-carbolines present in exudates ofO. tuberosa roots (Bais et al., 2002a), should greatly increase the opportunities available to study the molecular mechanisms underlying the secretion of phytochemicals.



ATP-Binding Cassette (ABC) Transporter as an Alternative to Vesicular Trafficking



The previous section highlighted the possibility of vesicular trafficking and fusion as a cellular mechanism responsible for root exudation, but could other mechanisms also be responsible once the compounds reach the membrane? For example, the involvement of membrane transporters such as the ABC transporters might be responsible for the secretion of root-secreted compounds. The ABC superfamily of membrane transporters is one of the largest protein families, and its members can be found in animals, bacteria, fungi, and plants. ABC transporters use ATP hydrolysis to actively transport chemically and structurally unrelated compounds from cells (Martinoia et al., 2002). The recent completion of the Arabidopsis genome research project (Arabidopsis Genome Initiative, 2000) revealed that Arabidopsis contains 53 putative ABC transporter genes. However, the protein localization and function of most of these genes are largely unknown (Martinoia et al., 2002). Most of the plant ABC transporters characterized to date have been localized in the vacuolar membrane and are believed to be responsible for the intracellular sequestration of cytotoxins (Theodoulou, 2000).

Currently, very little is known about plant plasma membrane ABC transporters, but the Arabidopsis AtPGP1, localized to the plasma membrane (Sidler et al., 1998), has been shown to be involved in cell elongation by actively pumping auxin from its site of synthesis in the cytoplasm to appropriate cells (Noh et al., 2001). Working on the assumption that plasma membrane ABC transporters might be involved in the secretion of defense metabolites, and their expression may be regulated by the concentration of these metabolites, Jasinski et al. (2002) identified a plasma membrane ABC transporter (NpABC1) from Nicotiana plumbaginifolia by treating cell cultures with various secondary metabolites. Interestingly, addition of sclareolide, an antifungal diterpene produced at the leaf surface of Nicotiana spp. (Baily et al., 1975), resulted in the expression of NpABC1 (Jasinski et al., 2002). These findings suggest that NpABC1 and likely other plasma membrane ABC transporters are involved in the secretion of secondary metabolites involved in plant defense, but further studies are required to positively identify plasma membrane ABC transporters involved in root exudation of specific compounds.





Previous SectionNext Section

Spatial Localization of Root Exudates



Major differences in root architecture exist among plant species (Fitter, 1996), and because different root classes of the same plant exploit different portions of the soil and are subject to different external signals, it has been speculated that they may have different metabolic activity. In accordance, it has been observed that nutrient influx by plant roots is heterogeneous in time and space. In the common bean (Phaseolus vulgaris), the basal roots have a consistently higher influx rate of nutrients than the other root classes (i.e. adventitious, lateral, and tap; Liao et al., 2001; Rubio et al., 2001). This characteristic could be beneficial for the plant because basal roots generally explore the topsoil, where the majority of available nutrients are located (Lynch and Brown, 2001). Furthermore, Russell and Sanderson (1967) found a large variation in the phosphorus influx rate among seminal, nodal, and lateral roots of barley (Hordeum vulgare).Kuhllmann and Barraclough (1987) observed that the rates of nitrogen uptake by nodal roots of wheat were up to 6 times higher than those of seminal roots, but the uptake ratio of potassium differed to a much smaller extent among root classes. Despite this large body of evidence linking root architecture with root absorption of nutrients, the effect of root architecture on root exudation has been virtually unexplored.

Another long-standing question is related to the pattern of root exudation along the longitudinal root axis. From the base to the tip, most root classes can be clearly divided into different sections based on marked dissimilarities in their anatomical characteristics (Gilroy and Jones, 2000). These sections are typically the root tip, the elongation zone, the maturation zone, and the matured zone. The root tip includes two subsections: the root cap and the meristematic region. In the elongation zone, located right behind the root tip, no cell division occurs, but there is vigorous cell elongation activity. The next section is the maturation zone, where xylem vessels are completely differentiated. Here, some epidermal cells elongate perpendicularly toward the rhizosphere; these cells are known as the root hairs. After a short period of life, root hairs die and this region becomes the mature zone of the root. The degree of cell vacuolization increases from the root tip (where no cell vacuoles are present) to the base of the root. How this anatomical heterogeneity along the root axis relates to the metabolic activity of the roots has concerned researchers for decades (Prevot and Steward, 1936).

Although the stages of aging correlate well with the metabolic activity of the root, it is widely recognized that the gradual maturation of root tissues along the root axis is not the only source of variation of metabolic activity (Eshel and Waisel, 1996). Although the large carbon demand in the apical zone has been traditionally attributed to high biosynthesis rates, it may also be due to an active root exudation process. In the case of the influx processes, the absorption of sulfur is highest in the elongation zone immediately behind the meristematic region (Holobrada, 1977) and that of iron at the apical zones of the roots. In the case of nitrogen or phosphorus, contrasting results have been found (Colmer and Bloom, 1998).

Much less attention has been focused on the spatial localization of the root exudation process. The scarce information available suggests that the pattern of exudation is not homogeneous along the root axis. Release of phytosiderophores in response to iron deficiencies appears to be concentrated in the apical zones of the root (Marschner et al., 1987). Release of organic anions would also follow a heterogeneous pattern along the root (Hoffland et al., 1989), which is consistent with the presence of a pH gradient from the tip to the base of the root (Fischer et al., 1989). On the other hand, based on the type of soil and its surface resistance, root tips may secrete a battery of compounds to soften the soil to facilitate root growth (Morel et al., 1991). Although such a mechanism has been hypothesized for decades, the chemicals involved in this phenomenon have yet to be identified. An understanding of the spatial and physical localization of the sites of exudation in the roots will facilitate the elucidation of plant-microbe and plant-plant interactions. For instance, external signals from pathogens and invasive plants may determine the zone of the root where the release of exudates takes place. If there is any relationship between the presence of pathogens and invasive plants with the localization of root exudation process, it is virtually unknown at the present time.



Previous SectionNext Section

FINAL REMARKS



Due to significant advances in root biology and current National Science Foundation-funded projects on genomics of root-specific traits, roots are no longer considered an unexplored biological frontier. In contrast, knowledge of rhizospheric processes mediated by root exudates has not developed at the same pace. As highlighted in this update, several lines of evidence indicate that root exudates in their various forms may regulate plant and microbial communities in the rhizosphere. It is worth mentioning that most microbes live in the soil, but just a few of these organisms have developed compatible interactions with specific plants to become successful plant pathogens. Instead, the vast majority of microbes exhibit incompatible interactions with plants, which could be explained by the constant and diverse secretion of antimicrobial root exudates. The understanding of the biology of root exudation processes may contribute to devising novel strategies for improving plant fitness and the isolation of novel value-added compounds found in the root exudates.



Previous SectionNext Section

Footnotes

  • ↵1 This work was supported by the Colorado State University Agricultural Experiment Station (to J.M.V.), by National Science Foundation-Faculty Early Career Development Award (CAREER) (grant no. MCB 0093014 to J.M.V.), by the Invasive Weeds Initiative of the State of Colorado (to J.M.V.), by the Lindbergh Foundation (to J.M.V.), by the Environmental Protection Agency (to J.M.V.), by the U.S. Department of Agriculture-National Research Initiative Competitive Grants Program (grant no. 2002–01267 to E.G.), and by the National Science Foundation (grant no. MCB 0130062 to E.G.).​
  • ↵* Corresponding author; email [email protected]; fax 970–491–7745.​
  • Received December 23, 2002.​
  • Revision received February 4, 2003.​
  • Accepted February 25, 2003.​
Previous Section




LITERATURE CITED

    • Arabidopsis Genome Initiative​
    (2000) Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature 408:796–815.


    CrossRefMedline

    • Baily JA,​
    • Carter GA,​
    • Burden RS,​
    • Wain RL​
    (1975) Control of rust diseases by diterpenes from Nicotiana glutinosa. Nature 255:328–329.


    CrossRef

    • Bais HP,​
    • Loyola Vargas VM,​
    • Flores HE,​
    • Vivanco JM​
    (2001) Root specific metabolism: the biology and biochemistry of underground organs. In vitro Cell Dev Biol Plant 37:730–741.


    CrossRefWeb of Science

    • Bais HP,​
    • Park S-W,​
    • Stermitz FR,​
    • Halligan KM,​
    • Vivanco JM​
    (2002a) Exudation of fluorescent β-carbolines from Oxalis tuberosa L. roots. Phytochemistry 61:539–543.


    CrossRefMedlineWeb of Science

    • Bais HP,​
    • Walker TS,​
    • Schweizer HP,​
    • Vivanco JM​
    (2002b) Root specific elicitation and antimicrobial activity of rosmarinic acid in hairy root cultures of sweet basil (Ocimum basilicum L.). Plant Physiol Biochem 40:983–995.




    • Bais HP,​
    • Walker TS,​
    • Stermitz FR,​
    • Hufbauer RA,​
    • Vivanco JM​
    (2002c) Enantiomeric dependent phytotoxic and antimicrobial activity of (±)-catechin; a rhizosecreted racemic mixture from Centaurea maculosa (spotted knapweed). Plant Physiol 128:1173–1179.


    FREE Full Text

    • Becard G,​
    • Douds DD,​
    • Pfeffer PE​
    (1992) Extensive in vitro hyphal growth of vesicular-arbuscular mycorrhizal fungi in presence of CO2 and flavonols. Appl Environ Microbiol 58:821–825.


    Abstract/FREE Full Text

    • Becard G,​
    • Taylor LP,​
    • Douds DD,​
    • Pfeffer PE,​
    • Doner LW​
    (1995) Flavonoids are not necessary plant signal compounds in arbuscular mycorrhizal symbiosis. Mol Plant-Microbe Interact 8:252–258.




    • Benfey PN,​
    • Scheres B​
    (2000) Root development. Curr Biol 16:R813–815.




    • Bengough AG,​
    • McKenzie BM​
    (1997) Sloughing of root cap cells decreases the frictional resistance to maize (Zea mays L.) root growth. J Exp Bot 48:885–893.




    • Brigham LA,​
    • Michaels PJ,​
    • Flores HE​
    (1999) Cell-specific production and antimicrobial activity of naphthoquinones in roots of Lithospermum erythrorhizon. Plant Physiol 119:417–428.


    Abstract/FREE Full Text

    • Callaway RM,​
    • Aschehoug ET​
    (2000) Invasive plants versus their new and old neighbors: a mechanism for exotic invasion. Science 90:521–523.




    • Chaboud A,​
    • Rougier M​
    (1984) Identification and localization of sugar components of rice (Oryza sativa L.) root cap mucilage. J Plant Physiol 116:323–330.




    • Colmer TD,​
    • Bloom AJ​
    (1998) A comparison of net NH4+ and NO3- fluxes along roots of rice and maize. Plant Cell Environ 21:240–246.


    CrossRef

    • Waisel Y,​
    • Eshel A,​
    • Kafkafi U​
    • Eshel A,​
    • Waisel Y​
    (1996) Multiform and multifunction of various constituents of one root system. in Plant Roots: The Hidden Half. eds WaiselY, EshelA, KafkafiU (Marcel Dekker, New York), pp 175–192.




    • Estabrook EM,​
    • Yoder JI​
    (1998) Plant-plant communications: rhizosphere signaling between parasitic angiosperms and their hosts. Plant Physiol 116:1–7.


    FREE Full Text

    • Facchini PJ​
    (2001) Alkaloid biosynthesis in plants: biochemistry, cell biology, molecular regulation, and metabolic engineering applications. Annu Rev Plant Physiol Plant Mol Biol 52:29–66.


    CrossRefMedlineWeb of Science

    • Fischer KS,​
    • Edmeades GO,​
    • Johnson EC​
    (1989) Selection for the improvement of maize yield under moisture-deficits. Field Crop Res 22:227–243.


    CrossRef

    • Waisel EAY,​
    • Kafkafi U​
    • Fitter A​
    (1996) Characteristics and functions of root systems. in Plant Roots: The Hidden Half. eds WaiselEAY, KafkafiU (Marcel Dekker, New York), pp 1–20.




    • Flores HE,​
    • Pickard JJ,​
    • Hoy MW​
    (1988) Production of polyacetylenes and thiophenes in heterotrophic and photosynthetic root cultures of Asteraceae. Biol Mol 7:233–254.




    • Flores HE,​
    • Vivanco JM,​
    • Loyola-Vargas VM​
    (1999) “Radicle” biochemistry: the biology of root-specific metabolism. Trends Plant Sci 4:220–226.


    CrossRefMedline

    • Fray RG​
    (2002) Altering plant-microbe interaction through artificially manipulating bacterial quorum sensing. Ann Bot 89:245–253.


    Abstract/FREE Full Text

    • Gilroy S,​
    • Jones DL​
    (2000) From form to function: development and nutrient uptake in root hairs. Trends Plant Sci 5:56–60.


    CrossRefMedlineWeb of Science

    • Griffin GJ,​
    • Hale MG,​
    • Shay FJ​
    (1976) Nature and quantity of sloughed organic matter produced by roots of axenic peanut plants. Soil Biol Biochem 8:29–32.


    CrossRef

    • Grotewold E​
    (2001) Subcellular trafficking of phytochemicals. Rec Res Dev Plant Physiol 2:31–48.




    • Grotewold E,​
    • Chamberlain M,​
    • St. Claire G,​
    • Swenson J,​
    • Siame BA,​
    • Butler LG,​
    • Snook M,​
    • Bowen B​
    (1998) Engineering secondary metabolism in maize cells by ectopic expression of transcription factors. Plant Cell 10:721–740.


    Abstract/FREE Full Text

    • Hawes MC,​
    • Gunawardena U,​
    • Miyasaka S,​
    • Zhao X​
    (2000) The role of root border cells in plant defense. Trends Plant Sci 5:128–133.


    CrossRefMedlineWeb of Science

    • Hoffland E,​
    • Findenegg GR,​
    • Nelemans JA​
    (1989) Solubilization of rock phosphate by rape. II. Local root exudation of organic acids as a response to P-starvation. Plant Soil 113:161–165.


    CrossRef

    • Holobrada M​
    (1977) Changes in sulphate uptake and accumulation along the primary root during tissue differentiation. Biol Plant 19:331–337.




    • Hutchison WD,​
    • Campbell CD​
    (1994) Economic impact of the sugarbeet root aphid (Homoptera: Aphididae) on sugarbeet yield and quality in southern Minn. J Econ Entomol 87:465–475.




    • Jasinski M,​
    • Stukkens Y,​
    • Degand H,​
    • Purnell B​
    (2002) A plant plasma membrane ATP binding cassette-type transporter is involved in antifungal terpenoid secretion. Plant Cell 13:1095–1107.




    • Keyes WJ,​
    • O'Malley RC,​
    • Kim D,​
    • Lynn DG​
    (2000) Signaling organogenesis in parasitic angiosperms: xenognosin generation, perception, and response. J Plant Growth Regul 19:217–231.


    Medline

    • Knee EM,​
    • Gong FC,​
    • Gao M,​
    • Teplitski M,​
    • Jones AR,​
    • Foxworthy A,​
    • Mort AJ,​
    • Bauer WD​
    (2001) Root mucilage from pea and its utilization by rhizosphere bacteria as a sole carbon source. Mol Plant-Microbe Interact 14:775–784.


    MedlineWeb of Science

    • Kuhllmann H,​
    • Barraclough PB​
    (1987) Comparison between the seminal and nodal root systems of winter wheat in their activity for N and K uptake. Z Pflanz Bodenkd 150:24–30.




    • Larson RA,​
    • Marley KA,​
    • Tuveson RW,​
    • Berenbaum MR​
    (1988) β-Carboline alkaloids: mechanisms of phototoxicity to bacteria and insects. Photochem Photobiol 48:665–674.


    Medline

    • Liao H,​
    • Rubio G,​
    • Yan X,​
    • Cao A,​
    • Brown KM,​
    • Lynch JP​
    (2001) Effect of phosphorus availability on basal root shallowness in the common bean. Plant Soil 232:69–79.


    CrossRefMedlineWeb of Science

    • Lynch JP,​
    • Brown KM​
    (2001) Topsoil foraging: an architectural adaptation of plants to low phosphorus availability. Plant Soil 237:225–237.


    CrossRef

    • Ma JF,​
    • Ryan PR,​
    • Delhaize E​
    (2001) Aluminum tolerance in plants and the complexing role of organic acids. Trends Plant Sci 6:273–278.


    CrossRefMedlineWeb of Science

    • Marschner H​
    (1995) Mineral Nutrition of Higher Plants (Academic Press, London), Ed 2.




    • Marschner H,​
    • Romheld V,​
    • Kissel M​
    (1987) Localization of phytosiderophore release and of iron uptake along intact barley roots. Physiol Plant 71:157–162.




    • Martinoia E,​
    • Klein M,​
    • Geisler M,​
    • Bovet L,​
    • Forestier C,​
    • Kolukisaoglu U,​
    • Muller-Rober B,​
    • Schulz B​
    (2002) Multifunctionality of plant ABC transporters: more than just detoxifiers. Planta 214:345–355.


    CrossRefMedlineWeb of Science

    • McCully ME​
    (1995) Water efflux from the surface of field-grown grass roots: observations of cryo-scanning electron microscopy. Physiol Plant 95:217–224.


    CrossRef

    • McCully ME,​
    • Boyer JS​
    (1997) The expansion of root cap mucilage during hydration: III. Changes in water potential and water content. Physiol Plant 99:169–177.




    • Morel J,​
    • Habib L,​
    • Plantureux S,​
    • Guckert A​
    (1991) Influence of maize root mucilage on soil aggregate stability. Plant Soil 136:111–119.




    • Nardi S,​
    • Concheri G,​
    • Pizzeghello D,​
    • Sturaro A,​
    • Rella R,​
    • Parvoli G​
    (2000) Soil organic matter mobilization by root exudates. Chemosphere 5:653–658.


    CrossRef

    • Noh B,​
    • Murphy AS,​
    • Spalding EP​
    (2001) Multidrug resistance-like genes of Arabidopsis required for auxin transport and auxin-mediated development. Plant Cell 13:2441–2454.


    Abstract/FREE Full Text

    • Peters NK,​
    • Frost JW,​
    • Long SR​
    (1986) A plant flavone, luteolin, induces expression of Rhizobium meliloti nodulation genes. Science 233:977–980.


    Abstract/FREE Full Text

    • Prevot P,​
    • Steward FC​
    (1936) Salient features of the root system relative to the problem of salt absorption. Plant Physiol 11:509–534.


    FREE Full Text

    • Ray TC,​
    • Callow JA,​
    • Kennedy JF​
    (1988) Composition of root mucilage polysaccharide from Lepidium sativum. J Exp Bot 39:1249–1261.


    Abstract/FREE Full Text

    • Tanner W,​
    • Loews FA​
    • Rougier M​
    (1981) Secretory activity at the root cap. in Encyclopedia of Plant Physiology, New Series, eds TannerW, LoewsFA (Springer Verlag, Berlin), 13B, Plant Carbohydrates II:542–574.




    • Rubio G,​
    • Walk T,​
    • Ge Z,​
    • Yan X,​
    • Liao H,​
    • Lynch JP​
    (2001) Root gravitropism and belowground competition among neighboring plants: a modeling approach. Ann Bot 88:929–940.


    Abstract/FREE Full Text

    • Russell RS,​
    • Sanderson J​
    (1967) Nutrient uptake by different parts of the intact roots of plants. J Exp Bot 18:491–508.


    Abstract/FREE Full Text

    • Ryan PR,​
    • Delhaize E​
    (2001) Function and mechanism of organic anion exudation from plant roots. Annu Rev Plant Physiol Mol Biol 52:527–560.


    CrossRefMedlineWeb of Science

    • Sidler M,​
    • Hassa P,​
    • Hasan S,​
    • Ringli C,​
    • Dudler R​
    (1998) Involvement of an ABC transporter in a developmental pathway regulating hypocotyls cell elongation in the light. Plant Cell 10:1623–1636.


    Abstract/FREE Full Text

    • Sims IM,​
    • Middleton K,​
    • Lane AG,​
    • Cairns AJ,​
    • Bacic A​
    (2000) Characterisation of extracellular polysaccharides from suspension cultures of members of the Poaceae. Planta 210:261–268.


    Medline

    • Smith DL,​
    • Fedoroff NV​
    (1995) LRP1, a gene expressed in lateral and adventitious root primordia of Arabidopsis. Plant Cell 7:735–745.


    Abstract/FREE Full Text

    • Snyder BA,​
    • Nicholson RL​
    (1990) Synthesis of phytoalexins in sorghum as a site-specific response to fungal ingress. Science 248:1637–1639.


    Abstract/FREE Full Text

    • Stintzi A,​
    • Browse J​
    (2000) The Arabidopsis male-sterile mutant, opr3, lacks the 12-oxophytodienoic acid reductase required for jasmonate synthesis. Proc Natl Acad Sci USA 97:10625–10630.


    Abstract/FREE Full Text

    • Stotz HU,​
    • Pittendrigh BR,​
    • Kroymann J,​
    • Weniger K,​
    • Fritsche J,​
    • Bauke A,​
    • Mitchell-Olds T​
    (2000) Induced plant defense responses against chewing insects. Ethylene signaling reduces resistance of Arabidopsis against Egyptian cotton worm but not diamondback moth. Plant Physiol 124:1007–1018.


    Abstract/FREE Full Text

    • Teplitski M,​
    • Robinson JB,​
    • Bauer WD​
    (2000) Plants secrete substances that mimic bacterial N-acyl homoserine lactone signal activities and affect population density-dependent behaviors in associated bacteria. Mol Plant-Microbe Interact 13:637–648.


    MedlineWeb of Science

    • Theodoulou FL​
    (2000) Plant ABC transporters. Biochem Biophys Acta 1465:79–103.


    Medline

    • Flores HE,​
    • Lynch JP,​
    • Eissentat D​
    • Trieu AT,​
    • Van Buuren ML,​
    • Harrison MJ​
    (1997) Gene expression in mycorrhizal roots of Medicago truncatula. in Radical Biology: Advances and Perspectives on the Function of Plant Roots. eds FloresHE, LynchJP, EissentatD (American Society of Plant Physiologists, Rockville, MD), pp 498–500.




    • Vermeer J,​
    • McCully ME​
    (1982) The rhizosheath of Zea: new insight into the structure and development. Planta 156:45–61.


    CrossRef

    • Verpoorte R,​
    • Alfermann AW​
    • Verpoorte R​
    (2000) Plant secondary metabolism. in Metabolic Engineering of Plant Secondary Metabolism. eds VerpoorteR, AlfermannAW (Kluwer Academic Publishers, Dordrecht, The Netherlands), pp 1–29.




    • Watt M,​
    • McCully ME,​
    • Canny MJ​
    (1994) Formation and stabilization of rhizosheaths of Zea mays L.: effect of soil water content. Plant Physiol 106:179–186.


    Abstract

    • Winkel-Shirley B​
    (2001) Flavonoid biosynthesis: a colorful model for genetics, biochemistry, cell biology and biotechnology. Plant Physiol 126:485–493.


    FREE Full Text

    • Wu T,​
    • Wittkamper J,​
    • Flores HE​
    (1999) Root herbivory in vitro: interaction between root and aphids grown in aseptic coculture. In Vitro Cell Dev Biol Plant 35:259–264.




    • Yoder JI​
    (2001) Host-plant recognition by parasitic Scrophulariaceae. Curr Opin Plant Biol 4:359–365.


    CrossRefMedlineWeb of Science

    • Young IM​
    (1995) Variation in moisture contents between bulk soil and the rhizosheath of Triticum aestivum L. cv. New Phytol 130:135–139.




Articles citing this article


  • P for Two, Sharing a Scarce Resource: Soil Phosphorus Acquisition in the Rhizosphere of Intercropped Species Plant Physiol. 2011 156: 1078-1086.

  • Amino Acid Export in Plants: A Missing Link in Nitrogen Cycling Mol Plant 2011 4: 453-463.

  • Mrt, a Gene Unique to Fungi, Encodes an Oligosaccharide Transporter and Facilitates Rhizosphere Competency in Metarhizium robertsii Plant Physiol. 2010 154: 1549-1557.

  • Influence of Forest Trees on the Distribution of Mineral Weathering-Associated Bacterial Communities of the Scleroderma citrinum Mycorrhizosphere Appl. Environ. Microbiol. 2010 76: 4780-4787.

  • Genetic and Metabolic Divergence within a Rhizobium leguminosarum bv. trifolii Population Recovered from Clover Nodules Appl. Environ. Microbiol. 2010 76: 4593-4600.

  • Regulation of flagellar, motility and chemotaxis genes in Rhizobium leguminosarum by the VisN/R-Rem cascade Microbiology 2010 156: 1673-1685.

  • Multiple adaptive responses of Australian native perennial legumes with pasture potential to grow in phosphorus- and moisture-limited environments Ann Bot 2010 105: 755-767.

  • Root Secretion of Phytochemicals in Arabidopsis Is Predominantly Not Influenced by Diurnal Rhythms Mol Plant 2010 3: 491-498.

  • Utricularia carnivory revisited: plants supply photosynthetic carbon to traps J Exp Bot 2010 61: 99-103.

  • Altered Profile of Secondary Metabolites in the Root Exudates of Arabidopsis ATP-Binding Cassette Transporter Mutants Plant Physiol. 2008 146: 762-771.

  • Metabolomics Reveals Novel Pathways and Differential Mechanistic and Elicitor-Specific Responses in Phenylpropanoid and Isoflavonoid Biosynthesis in Medicago truncatula Cell Cultures Plant Physiol. 2008 146: 387-402.

  • Progressive Inhibition by Water Deficit of Cell Wall Extensibility and Growth along the Elongation Zone of Maize Roots Is Related to Increased Lignin Metabolism and Progressive Stelar Accumulation of Wall Phenolics Plant Physiol. 2006 140: 603-612.

  • Paenibacillus polymyxa Invades Plant Roots and Forms Biofilms Appl. Environ. Microbiol. 2005 71: 7292-7300.

  • Structural Complexity, Differential Response to Infection, and Tissue Specificity of Indolic and Phenylpropanoid Secondary Metabolism in Arabidopsis Roots Plant Physiol. 2005 138: 1058-1070.

  • Detection of and Response to Signals Involved in Host-Microbe Interactions by Plant-Associated Bacteria Microbiol. Mol. Biol. Rev. 2005 69: 155-194.

  • Enhanced Mineralization of [U-14C]2,4-Dichlorophenoxyacetic Acid in Soil from the Rhizosphere of Trifolium pratense Appl. Environ. Microbiol. 2004 70: 4766-4774.

  • Characterization of a Root-Specific Arabidopsis Terpene Synthase Responsible for the Formation of the Volatile Monoterpene 1,8-Cineole Plant Physiol. 2004 135: 1956-1966.

 

billy4479

Moderator
holy shit man Root-Root Communication this is like porn for me jk but some good shit talks about them talking to the microbes and knowing when ather plants is growing next to them makes sence whe you think about it ...
 

cannawizard

Well-Known Member
holy shit man Root-Root Communication this is like porn for me jk but some good shit talks about them talking to the microbes and knowing when ather plants is growing next to them makes sence whe you think about it ...
*..man.. are we even on Cannabis anymore.. hahaha.. ahead of the curve ;) cheers
 
Top