Chemotaxis

Last updated

Capillary tube assay for chemotaxis. Motile prokaryotes sense chemicals in their environment and change their motility accordingly. Absent chemicals, movement is completely random. When an attractant or repellent is present, runs become longer and tumbles become less frequent. The result is net movement towards or away from the chemical (i.e., up or down the chemical gradient). The net movement can be seen in the beaker, where the bacteria accumulate around the origin of the attractant, and away from the origin of the repellent. Chemotaxis1.jpg
Capillary tube assay for chemotaxis. Motile prokaryotes sense chemicals in their environment and change their motility accordingly. Absent chemicals, movement is completely random. When an attractant or repellent is present, runs become longer and tumbles become less frequent. The result is net movement towards or away from the chemical (i.e., up or down the chemical gradient). The net movement can be seen in the beaker, where the bacteria accumulate around the origin of the attractant, and away from the origin of the repellent.

Chemotaxis (from chemo- + taxis ) is the movement of an organism or entity in response to a chemical stimulus. [1] Somatic cells, bacteria, and other single-cell or multicellular organisms direct their movements according to certain chemicals in their environment. This is important for bacteria to find food (e.g., glucose) by swimming toward the highest concentration of food molecules, or to flee from poisons (e.g., phenol). In multicellular organisms, chemotaxis is critical to early development (e.g., movement of sperm towards the egg during fertilization) and development (e.g., migration of neurons or lymphocytes) as well as in normal function and health (e.g., migration of leukocytes during injury or infection). [2] In addition, it has been recognized that mechanisms that allow chemotaxis in animals can be subverted during cancer metastasis. [3] The aberrant chemotaxis of leukocytes and lymphocytes also contribute to inflammatory diseases such as atherosclerosis, asthma, and arthritis. [4] [5] [6] [7] Sub-cellular components, such as the polarity patch generated by mating yeast, may also display chemotactic behavior. [8]

Contents

Positive chemotaxis occurs if the movement is toward a higher concentration of the chemical in question; negative chemotaxis if the movement is in the opposite direction. Chemically prompted kinesis (randomly directed or nondirectional) can be called chemokinesis.

History of chemotaxis research

Although migration of cells was detected from the early days of the development of microscopy by Leeuwenhoek, a Caltech lecture regarding chemotaxis propounds that 'erudite description of chemotaxis was only first made by T. W. Engelmann (1881) and W. F. Pfeffer (1884) in bacteria, and H. S. Jennings (1906) in ciliates'. [9] The Nobel Prize laureate I. Metchnikoff also contributed to the study of the field during 1882 to 1886, with investigations of the process as an initial step of phagocytosis. [10] The significance of chemotaxis in biology and clinical pathology was widely accepted in the 1930s, and the most fundamental definitions underlying the phenomenon were drafted by this time.[ by whom? ] The most important aspects in quality control of chemotaxis assays were described by H. Harris in the 1950s. [11] In the 1960s and 1970s, the revolution of modern cell biology and biochemistry provided a series of novel techniques that became available to investigate the migratory responder cells and subcellular fractions responsible for chemotactic activity. [12] The availability of this technology led to the discovery of C5a, a major chemotactic factor involved in acute inflammation. The pioneering works of J. Adler modernized Pfeffer's capillary assay and represented a significant turning point in understanding the whole process of intracellular signal transduction of bacteria. [13] [14]

Bacterial chemotaxis—general characteristics

Correlation of swimming behaviour and flagellar rotation ChtxCCW CW (Fixed).png
Correlation of swimming behaviour and flagellar rotation

Some bacteria, such as E. coli , have several flagella per cell (4–10 typically). These can rotate in two ways:

  1. Counter-clockwise rotation aligns the flagella into a single rotating bundle, causing the bacterium to swim in a straight line; and
  2. Clockwise rotation breaks the flagella bundle apart such that each flagellum points in a different direction, causing the bacterium to tumble in place. [15]

The directions of rotation are given for an observer outside the cell looking down the flagella toward the cell. [16]

Behavior

The overall movement of a bacterium is the result of alternating tumble and swim phases, called run-and-tumble motion. [17] As a result, the trajectory of a bacterium swimming in a uniform environment will form a random walk with relatively straight swims interrupted by random tumbles that reorient the bacterium. [18] Bacteria such as E. coli are unable to choose the direction in which they swim, and are unable to swim in a straight line for more than a few seconds due to rotational diffusion; in other words, bacteria "forget" the direction in which they are going. By repeatedly evaluating their course, and adjusting if they are moving in the wrong direction, bacteria can direct their random walk motion toward favorable locations. [19]

In the presence of a chemical gradient bacteria will chemotax, or direct their overall motion based on the gradient. If the bacterium senses that it is moving in the correct direction (toward attractant/away from repellent), it will keep swimming in a straight line for a longer time before tumbling; however, if it is moving in the wrong direction, it will tumble sooner. Bacteria like E. coli use temporal sensing to decide whether their situation is improving or not, and in this way, find the location with the highest concentration of attractant, detecting even small differences in concentration. [20]

This biased random walk is a result of simply choosing between two methods of random movement; namely tumbling and straight swimming. [21] The helical nature of the individual flagellar filament is critical for this movement to occur. The protein structure that makes up the flagellar filament, flagellin, is conserved among all flagellated bacteria. [22] Vertebrates seem to have taken advantage of this fact by possessing an immune receptor (TLR5) designed to recognize this conserved protein. [23]

As in many instances in biology, there are bacteria that do not follow this rule. Many bacteria, such as Vibrio, are monoflagellated and have a single flagellum at one pole of the cell. Their method of chemotaxis is different. Others possess a single flagellum that is kept inside the cell wall. These bacteria move by spinning the whole cell, which is shaped like a corkscrew. [24] [ page needed ]

Signal transduction

Domain structure of chemotaxis receptor for Asp ChtxAspRec.png
Domain structure of chemotaxis receptor for Asp

Chemical gradients are sensed through multiple transmembrane receptors, called methyl-accepting chemotaxis proteins (MCPs), which vary in the molecules that they detect. [25] Thousands of MCP receptors are known to be encoded across the bacterial kingdom. [26] These receptors may bind attractants or repellents directly or indirectly through interaction with proteins of periplasmatic space. [27] The signals from these receptors are transmitted across the plasma membrane into the cytosol, where Che proteins are activated. [28] The Che proteins alter the tumbling frequency, and alter the receptors. [28]

Flagellum regulation

The proteins CheW and CheA bind to the receptor. The absence of receptor activation results in autophosphorylation in the histidine kinase, CheA, at a single highly conserved histidine residue. [29] [ better source needed ] CheA, in turn, transfers phosphoryl groups to conserved aspartate residues in the response regulators CheB and CheY; CheA is a histidine kinase and it does not actively transfer the phosphoryl group, rather, the response regulator CheB takes the phosphoryl group from CheA.[ citation needed ] This mechanism of signal transduction is called a two-component system, and it is a common form of signal transduction in bacteria.[ citation needed ] CheY induces tumbling by interacting with the flagellar switch protein FliM, inducing a change from counter-clockwise to clockwise rotation of the flagellum. Change in the rotation state of a single flagellum can disrupt the entire flagella bundle and cause a tumble.[ citation needed ]

Receptor regulation

Signalling pathways of E.coli Chtxbactsign1.png
Signalling pathways of E.coli

CheB, when activated by CheA, acts as a methylesterase, removing methyl groups from glutamate residues on the cytosolic side of the receptor; it works antagonistically with CheR, a methyltransferase, which adds methyl residues to the same glutamate residues. [30] If the level of an attractant remains high, the level of phosphorylation of CheA (and, therefore, CheY and CheB) will remain low, the cell will swim smoothly, and the level of methylation of the MCPs will increase (because CheB-P is not present to demethylate). [30] The MCPs no longer respond to the attractant when they are fully methylated; therefore, even though the level of attractant might remain high, the level of CheA-P (and CheB-P) increases and the cell begins to tumble. [30] The MCPs can be demethylated by CheB-P, and, when this happens, the receptors can once again respond to attractants. [30] The situation is the opposite with regard to repellents: fully methylated MCPs respond best to repellents, while least-methylated MCPs respond worst to repellents.[ citation needed ] This regulation allows the bacterium to 'remember' chemical concentrations from the recent past, a few seconds, and compare them to those it is currently experiencing, thus 'know' whether it is traveling up or down a gradient. [31] that bacteria have to chemical gradients, other mechanisms are involved in increasing the absolute value of the sensitivity on a given background. Well-established examples are the ultra-sensitive response of the motor to the CheY-P signal, and the clustering of chemoreceptors. [32] [33]

Chemoattractants and chemorepellents

Chemoattractants and chemorepellents are inorganic or organic substances possessing chemotaxis-inducer effect in motile cells. These chemotactic ligands create chemical concentration gradients that organisms, prokaryotic and eukaryotic, move toward or away from, respectively. [34]

float Chtx-AttrRep-en.png
float

Effects of chemoattractants are elicited via chemoreceptors such as methyl-accepting chemotaxis proteins (MCP). [35] MCPs in E.coli include Tar, Tsr, Trg and Tap. [36] Chemoattracttants to Trg include ribose and galactose with phenol as a chemorepellent. Tap and Tsr recognize dipeptides and serine as chemoattractants, respectively. [36]

Chemoattractants or chemorepellents bind MCPs at its extracellular domain; an intracellular signaling domain relays the changes in concentration of these chemotactic ligands to downstream proteins like that of CheA which then relays this signal to flagellar motors via phosphorylated CheY (CheY-P). [35] CheY-P can then control flagellar rotation influencing the direction of cell motility. [35]

For E.coli, S. meliloti , and R. spheroides, the binding of chemoattractants to MCPs inhibit CheA and therefore CheY-P activity, resulting in smooth runs, but for B. substilis , CheA activity increases. [35] Methylation events in E.coli cause MCPs to have lower affinity to chemoattractants which causes increased activity of CheA and CheY-P resulting in tumbles. [35] In this way cells are able to adapt to the immediate chemoattractant concentration and detect further changes to modulate cell motility. [35]

Chemoattractants in eukaryotes are well characterized for immune cells. Formyl peptides, such as fMLF, attract leukocytes such as neutrophils and macrophages, causing movement toward infection sites. [37] Non-acylated methioninyl peptides do not act as chemoattractants to neutrophils and macrophages. [37] Leukocytes also move toward chemoattractants C5a, a complement component, and pathogen-specific ligands on bacteria. [37]

Mechanisms concerning chemorepellents are less known than chemoattractants. Although chemorepellents work to confer an avoidance response in organisms, Tetrahymena thermophila adapt to a chemorepellent, Netrin-1 peptide, within 10 minutes of exposure; however, exposure to chemorepellents such as GTP, PACAP-38, and nociceptin show no such adaptations. [38] GTP and ATP are chemorepellents in micro-molar concentrations to both Tetrahymena and Paramecium . These organisms avoid these molecules by producing avoiding reactions to re-orient themselves away from the gradient. [39]

Eukaryotic chemotaxis

Difference of gradient sensing in prokaryotes and eukaryotes Chtxbaceukkl1.png
Difference of gradient sensing in prokaryotes and eukaryotes

The mechanism of chemotaxis that eukaryotic cells employ is quite different from that in the bacteria E. coli; however, sensing of chemical gradients is still a crucial step in the process. [40] [ better source needed ] Due to their small size and other biophysical constraints, E. coli cannot directly detect a concentration gradient. [41] Instead, they employ temporal gradient sensing, where they move over larger distances several times their own width and measure the rate at which perceived chemical concentration changes. [42] [43]

Eukaryotic cells are much larger than prokaryotes and have receptors embedded uniformly throughout the cell membrane. [42] Eukaryotic chemotaxis involves detecting a concentration gradient spatially by comparing the asymmetric activation of these receptors at the different ends of the cell. [42] Activation of these receptors results in migration towards chemoattractants, or away from chemorepellants. [42] In mating yeast, which are non-motile, patches of polarity proteins on the cell cortex can relocate in a chemotactic fashion up pheromone gradients. [44] [8]

It has also been shown that both prokaryotic and eukaryotic cells are capable of chemotactic memory. [43] [45] In prokaryotes, this mechanism involves the methylation of receptors called methyl-accepting chemotaxis proteins (MCPs). [43] This results in their desensitization and allows prokaryotes to "remember" and adapt to a chemical gradient. [43] In contrast, chemotactic memory in eukaryotes can be explained by the Local Excitation Global Inhibition (LEGI) model. [45] [46] LEGI involves the balance between a fast excitation and delayed inhibition which controls downstream signaling such as Ras activation and PIP3 production. [47]

Levels of receptors, intracellular signalling pathways and the effector mechanisms all represent diverse, eukaryotic-type components. In eukaryotic unicellular cells, amoeboid movement and cilium or the eukaryotic flagellum are the main effectors (e.g., Amoeba or Tetrahymena). [48] [49] Some eukaryotic cells of higher vertebrate origin, such as immune cells also move to where they need to be. Besides immune competent cells (granulocyte, monocyte, lymphocyte) a large group of cells—considered previously to be fixed into tissues—are also motile in special physiological (e.g., mast cell, fibroblast, endothelial cells) or pathological conditions (e.g., metastases). [50] Chemotaxis has high significance in the early phases of embryogenesis as development of germ layers is guided by gradients of signal molecules. [51] [52]

Motility

Unlike motility in bacterial chemotaxis, the mechanism by which eukaryotic cells physically move is unclear. There appear to be mechanisms by which an external chemotactic gradient is sensed and turned into an intracellular PIP3 gradient, which results in a gradient and the activation of a signaling pathway, culminating in the polymerisation of actin filaments. The growing distal end of actin filaments develops connections with the internal surface of the plasma membrane via different sets of peptides and results in the formation of anterior pseudopods and posterior uropods. Cilia of eukaryotic cells can also produce chemotaxis; in this case, it is mainly a Ca2+-dependent induction of the microtubular system of the basal body and the beat of the 9 + 2 microtubules within cilia. The orchestrated beating of hundreds of cilia is synchronized by a submembranous system built between basal bodies. The details of the signaling pathways are still not totally clear.

Chemotaxis related migratory responses Chtxphenomen1.png
Chemotaxis related migratory responses

Chemotaxis refers to the directional migration of cells in response to chemical gradients; several variations of chemical-induced migration exist as listed below.

  • Chemokinesis refers to an increase in cellular motility in response to chemicals in the surrounding environment. Unlike chemotaxis, the migration stimulated by chemokinesis lacks directionality, and instead increases environmental scanning behaviors. [53]
  • In haptotaxis the gradient of the chemoattractant is expressed or bound on a surface, in contrast to the classical model of chemotaxis, in which the gradient develops in a soluble fluid. [54] The most common biologically active haptotactic surface is the extracellular matrix (ECM); the presence of bound ligands is responsible for induction of transendothelial migration and angiogenesis.
  • Necrotaxis embodies a special type of chemotaxis when the chemoattractant molecules are released from necrotic or apoptotic cells. Depending on the chemical character of released substances, necrotaxis can accumulate or repel cells, which underlines the pathophysiological significance of this phenomenon.

Receptors

In general, eukaryotic cells sense the presence of chemotactic stimuli through the use of 7-transmembrane (or serpentine) heterotrimeric G-protein-coupled receptors, a class representing a significant portion of the genome. [55] Some members of this gene superfamily are used in eyesight (rhodopsins) as well as in olfaction (smelling). [56] [57] The main classes of chemotaxis receptors are triggered by:

However, induction of a wide set of membrane receptors (e.g., cyclic nucleotides, amino acids, insulin, vasoactive peptides) also elicit migration of the cell. [59]

Chemotactic selection

Chemotactic selection Chtxsel.png
Chemotactic selection

While some chemotaxis receptors are expressed in the surface membrane with long-term characteristics, as they are determined genetically, others have short-term dynamics, as they are assembled ad hoc in the presence of the ligand. [60] The diverse features of the chemotaxis receptors and ligands allows for the possibility of selecting chemotactic responder cells with a simple chemotaxis assay By chemotactic selection, we can determine whether a still-uncharacterized molecule acts via the long- or the short-term receptor pathway. [61] The term chemotactic selection is also used to designate a technique that separates eukaryotic or prokaryotic cells according to their chemotactic responsiveness to selector ligands. [62] [ non-primary source needed ][ non-primary source needed ]

Chemotactic ligands

Structure of chemokine classes ChtxChemokineStruct.png
Structure of chemokine classes
Three dimensional structure of chemokines ChtxChemkinStr2.png
Three dimensional structure of chemokines

The number of molecules capable of eliciting chemotactic responses is relatively high, and we can distinguish primary and secondary chemotactic molecules.[ citation needed ] The main groups of the primary ligands are as follows:

Chemotactic range fitting

Chemotactic range fitting ChtxCRF2.png
Chemotactic range fitting

Chemotactic responses elicited by ligand-receptor interactions vary with the concentration of the ligand. Investigations of ligand families (e.g. amino acids or oligopeptides) demonstrates that chemoattractant activity occurs over a wide range, while chemorepellent activities have narrow ranges. [69]

Clinical significance

A changed migratory potential of cells has relatively high importance in the development of several clinical symptoms and syndromes. Altered chemotactic activity of extracellular (e.g., Escherichia coli) or intracellular (e.g., Listeria monocytogenes) pathogens itself represents a significant clinical target. Modification of endogenous chemotactic ability of these microorganisms by pharmaceutical agents can decrease or inhibit the ratio of infections or spreading of infectious diseases. Apart from infections, there are some other diseases wherein impaired chemotaxis is the primary etiological factor, as in Chédiak–Higashi syndrome, where giant intracellular vesicles inhibit normal migration of cells.

Chemotaxis in diseases[ citation needed ]
Type of diseaseChemotaxis increasedChemotaxis decreased
Infections Inflammations AIDS, Brucellosis
Chemotaxis results in the disease Chédiak–Higashi syndrome, Kartagener syndrome
Chemotaxis is affected Atherosclerosis, arthritis, periodontitis, psoriasis, reperfusion injury, metastatic tumors Multiple sclerosis, Hodgkin disease, male infertility
Intoxications Asbestos, benzpyrene Hg and Cr salts, ozone

Mathematical models

Several mathematical models of chemotaxis were developed depending on the type of

Although interactions of the factors listed above make the behavior of the solutions of mathematical models of chemotaxis rather complex, it is possible to describe the basic phenomenon of chemotaxis-driven motion in a straightforward way. Indeed, let us denote with the spatially non-uniform concentration of the chemo-attractant and as its gradient. Then the chemotactic cellular flow (also called current) that is generated by the chemotaxis is linked to the above gradient by the law: [70]

where is the spatial density of the cells and is the so-called 'Chemotactic coefficient' - is often not constant, but a decreasing function of the chemo-attractant. For some quantity that is subject to total flux and generation/destruction term , it is possible to formulate a continuity equation:

where is the divergence. This general equation applies to both the cell density and the chemo-attractant. Therefore, incorporating a diffusion flux into the total flux term, the interactions between these quantities are governed by a set of coupled reaction-diffusion partial differential equations describing the change in and : [70]

where describes the growth in cell density, is the kinetics/source term for the chemo-attractant, and the diffusion coefficients for cell density and the chemo-attractant are respectively and .

Spatial ecology of soil microorganisms is a function of their chemotactic sensitivities towards substrate and fellow organisms. [71] [ non-primary source needed ][ non-primary source needed ] The chemotactic behavior of the bacteria was proven to lead to non-trivial population patterns even in the absence of environmental heterogeneities. The presence of structural pore scale heterogeneities has an extra impact on the emerging bacterial patterns.

Measurement of chemotaxis

A wide range of techniques is available to evaluate chemotactic activity of cells or the chemoattractant and chemorepellent character of ligands. The basic requirements of the measurement are as follows:

Despite the fact that an ideal chemotaxis assay is still not available, there are several protocols and pieces of equipment that offer good correspondence with the conditions described above. The most commonly used are summarised in the table below:

Type of assayAgar-plate assaysTwo-chamber assaysOthers
Examples
  • PP-chamber
  • Boyden chamber
  • Zigmond chamber
  • Dunn chambers
  • Multi-well chambers
  • Capillary techniques
  • T-maze technique
  • Opalescence technique
  • Orientation assays

Artificial chemotactic systems

Chemical robots that use artificial chemotaxis to navigate autonomously have been designed. [72] [73] Applications include targeted delivery of drugs in the body. [74] More recently, enzyme molecules have also shown positive chemotactic behavior in the gradient of their substrates. [75] The thermodynamically-favorable binding of enzymes to their specific substrates is recognized as the origin of enzymatic chemotaxis. [76] Additionally, enzymes in cascades have also shown substrate-driven chemotactic aggregation. [77]

Apart from active enzymes, non-reacting molecules also show chemotactic behavior. This has been demonstrated by using dye molecules that move directionally in gradients of polymer solution through favorable hydrophobic interactions. [78]

See also

Related Research Articles

A taxis is the movement of an organism in response to a stimulus such as light or the presence of food. Taxes are innate behavioural responses. A taxis differs from a tropism in that in the case of taxis, the organism has motility and demonstrates guided movement towards or away from the stimulus source. It is sometimes distinguished from a kinesis, a non-directional change in activity in response to a stimulus.

Chemotropism is defined as the growth of organisms navigated by chemical stimulus from outside of the organism. It has been observed in bacteria, plants and fungi. A chemical gradient can influence the growth of the organism in a positive or negative way. Positive growth is characterized by growing towards a stimulus and negative growth is growing away from the stimulus.

In biology, cell signaling is the process by which a cell interacts with itself, other cells, and the environment. Cell signaling is a fundamental property of all cellular life in prokaryotes and eukaryotes.

Julius Adler is an American biochemist. He has been an Emeritus Professor of biochemistry and genetics at the University of Wisconsin–Madison since 1997.

<span class="mw-page-title-main">CCL7</span> Mammalian protein found in Homo sapiens

Chemokine ligand 7 (CCL7) is a small cytokine that was previously called monocyte-chemotactic protein 3 (MCP3). CCL7 is a small protein that belongs to the CC chemokine family and is most closely related to CCL2.

<span class="mw-page-title-main">Protein-glutamate O-methyltransferase</span>

In enzymology, a protein-glutamate O-methyltransferase is an enzyme that catalyzes the chemical reaction

<i>N</i>-Formylmethionine-leucyl-phenylalanine Chemical compound

N-Formylmethionyl-leucyl-phenylalanine is an N-formylated tripeptide and sometimes simply referred to as chemotactic peptide is a potent polymorphonuclear leukocyte (PMN) chemotactic factor and is also a macrophage activator.

<span class="mw-page-title-main">Protein-glutamate methylesterase</span>

The enzyme protein-glutamate methylesterase (EC 3.1.1.61) catalyzes the reaction

<span class="mw-page-title-main">Two-component regulatory system</span> Method of stimulus sensing and response in cells

In molecular biology, a two-component regulatory system serves as a basic stimulus-response coupling mechanism to allow organisms to sense and respond to changes in many different environmental conditions. Two-component systems typically consist of a membrane-bound histidine kinase that senses a specific environmental stimulus, and a corresponding response regulator that mediates the cellular response, mostly through differential expression of target genes. Although two-component signaling systems are found in all domains of life, they are most common by far in bacteria, particularly in Gram-negative and cyanobacteria; both histidine kinases and response regulators are among the largest gene families in bacteria. They are much less common in archaea and eukaryotes; although they do appear in yeasts, filamentous fungi, and slime molds, and are common in plants, two-component systems have been described as "conspicuously absent" from animals.

<span class="mw-page-title-main">Bacterial motility</span> Ability of bacteria to move independently using metabolic energy

Bacterial motility is the ability of bacteria to move independently using metabolic energy. Most motility mechanisms that evolved among bacteria also evolved in parallel among the archaea. Most rod-shaped bacteria can move using their own power, which allows colonization of new environments and discovery of new resources for survival. Bacterial movement depends not only on the characteristics of the medium, but also on the use of different appendages to propel. Swarming and swimming movements are both powered by rotating flagella. Whereas swarming is a multicellular 2D movement over a surface and requires the presence of surfactants, swimming is movement of individual cells in liquid environments.

<span class="mw-page-title-main">Phototaxis</span>

Phototaxis is a kind of taxis, or locomotory movement, that occurs when a whole organism moves towards or away from a stimulus of light. This is advantageous for phototrophic organisms as they can orient themselves most efficiently to receive light for photosynthesis. Phototaxis is called positive if the movement is in the direction of increasing light intensity and negative if the direction is opposite.

Sperm guidance is the process by which sperm cells (spermatozoa) are directed to the oocyte (egg) for the aim of fertilization. In the case of marine invertebrates the guidance is done by chemotaxis. In the case of mammals, it appears to be done by chemotaxis, thermotaxis and rheotaxis.

Sperm chemotaxis is a form of sperm guidance, in which sperm cells (spermatozoa) follow a concentration gradient of a chemoattractant secreted from the oocyte and thereby reach the oocyte.

<span class="mw-page-title-main">Chemotactic drug-targeting</span>

Targeted drug delivery is one of many ways researchers seek to improve drug delivery systems' overall efficacy, safety, and delivery. Within this medical field is a special reversal form of drug delivery called chemotactic drug targeting. By using chemical agents to help guide a drug carrier to a specific location within the body, this innovative approach seeks to improve precision and control during the drug delivery process, decrease the risk of toxicity, and potentially lower the required medical dosage needed. The general components of the conjugates are designed as follows: (i) carrier – regularly possessing promoter effect also on internalization into the cell; (ii) chemotactically active ligands acting on the target cells; (iii) drug to be delivered in a selective way and (iv) spacer sequence which joins drug molecule to the carrier and due to it enzyme labile moiety makes possible the intracellular compartment specific release of the drug. Careful selection of chemotactic component of the ligand not only the chemoattractant character could be expended, however, chemorepellent ligands are also valuable as they are useful to keep away cell populations degrading the conjugate containing the drug. In a larger sense, chemotactic drug-targeting has the potential to improve cancer, inflammation, and arthritis treatment by taking advantage of the difference in environment between the target site and its surroundings. Therefore, this Wikipedia article aims to provide a brief overview of chemotactic drug targeting, the principles behind the approach, possible limitations and advantages, and its application to cancer and inflammation.

<span class="mw-page-title-main">Methyl-accepting chemotaxis proteins</span> Family of bacterial transmembrane receptors

The methyl-accepting chemotaxis proteins are a family of transmembrane receptors that mediate chemotactic response in certain enteric bacteria, such as Salmonella enterica enterica and Escherichia coli. These methyl-accepting chemotaxis receptors are one of the first components in the sensory excitation and adaptation responses in bacteria, which act to alter swimming behaviour upon detection of specific chemicals. Use of the MCP allows bacteria to detect concentrations of molecules in the extracellular matrix so that the bacteria may smooth swim or tumble accordingly. If the bacterium detects rising levels of attractants (nutrients) or declining levels of repellents (toxins), the bacterium will continue swimming forward, or smooth swimming. If the bacterium detects declining levels of attractants or rising levels of repellents, the bacterium will tumble and re-orient itself in a new direction. In this manner, a bacterium may swim towards nutrients and away from toxins

Chemorepulsion is the directional movement of a cell away from a substance. Of the two directional varieties of chemotaxis, chemoattraction has been studied to a much greater extent. Only recently have the key components of the chemorepulsive pathway been elucidated. The exact mechanism is still being investigated, and its constituents are currently being explored as likely candidates for immunotherapies.

<span class="mw-page-title-main">Biomolecular gradient</span>

A biomolecular gradient is established by a difference in the concentration of molecules in a biological system such as individual cells, groups of cells, or an entire organism. A biomolecular gradient can exist intracellularly or extracellularly. The purposes of such gradients in biological systems vary, but include chemotaxis and functions in development. These types of gradients play a role in many different types of signaling as well as recently being implicated in cancer metastasis.

Robert Insall is the professor of computational cell biology at University College London and the University of Glasgow. His work focuses on how eukaryotic cells move, and how they choose the direction in which they move. He is known for demonstrating that cells can spread in the body and find their way through mazes by creating gradients of chemoattractants.

Michael Eisenbach is an Israeli biochemist who specializes in the navigation mechanisms of bacterial and sperm cells. He is a professor emeritus at the Weizmann Institute of Science, Department of Biomolecular Sciences, Rehovot, Israel. He discovered that sperm cells (spermatozoa) of mammals are actively guided to the egg. This opened the research field of mammalian sperm navigation. He demonstrated that the active navigation entails chemotaxis and thermotaxis. He made seminal contributions to the understanding of these two processes at the molecular, physiological and behavioural levels, as well as contributing to our understanding of the molecular mechanism of bacterial chemotaxis.

<span class="mw-page-title-main">Run-and-tumble motion</span> Type of bacterial motion

Run-and-tumble motion is a movement pattern exhibited by certain bacteria and other microscopic agents. It consists of an alternating sequence of "runs" and "tumbles": during a run, the agent propels itself in a fixed direction, and during a tumble, it remains stationary while it reorients itself in preparation for the next run.

References

  1. Chisholm, Hugh, ed. (1911). "Chemotaxis"  . Encyclopædia Britannica . Vol. 6 (11th ed.). Cambridge University Press. p. 77.
  2. de Oliveira S, Rosowski EE, Huttenlocher A (May 2016). "Neutrophil migration in infection and wound repair: going forward in reverse". Nature Reviews. Immunology. 16 (6): 378–91. doi:10.1038/nri.2016.49. PMC   5367630 . PMID   27231052.
  3. Stuelten CH, Parent CA, Montell DJ (May 2018). "Cell motility in cancer invasion and metastasis: insights from simple model organisms". Nature Reviews. Cancer. 18 (5): 296–312. doi:10.1038/nrc.2018.15. PMC   6790333 . PMID   29546880.
  4. Li J, Ley K (January 2015). "Lymphocyte migration into atherosclerotic plaque". Arteriosclerosis, Thrombosis, and Vascular Biology. 35 (1): 40–9. doi:10.1161/ATVBAHA.114.303227. PMC   4429868 . PMID   25301842.
  5. Gelfand EW (October 2017). "Importance of the leukotriene B4-BLT1 and LTB4-BLT2 pathways in asthma". Seminars in Immunology. 33: 44–51. doi:10.1016/j.smim.2017.08.005. PMC   5679233 . PMID   29042028.
  6. Planagumà A, Domènech T, Pont M, Calama E, García-González V, López R, Aulí M, López M, Fonquerna S, Ramos I, de Alba J, Nueda A, Prats N, Segarra V, Miralpeix M, Lehner MD (October 2015). "Combined anti CXC receptors 1 and 2 therapy is a promising anti-inflammatory treatment for respiratory diseases by reducing neutrophil migration and activation". Pulmonary Pharmacology & Therapeutics. 34: 37–45. doi:10.1016/j.pupt.2015.08.002. PMID   26271598.
  7. Rana AK, Li Y, Dang Q, Yang F (December 2018). "Monocytes in rheumatoid arthritis: Circulating precursors of macrophages and osteoclasts and, their heterogeneity and plasticity role in RA pathogenesis". International Immunopharmacology. 65: 348–359. doi:10.1016/j.intimp.2018.10.016. PMID   30366278. S2CID   53116963.
  8. 1 2 Ghose, Debraj; Jacobs, Katherine; Ramirez, Samuel; Elston, Timothy; Lew, Daniel (1 June 2021). "Chemotactic movement of a polarity site enables yeast cells to find their mates". Proceedings of the National Academy of Sciences. 118 (22): e2025445118. Bibcode:2021PNAS..11825445G. doi: 10.1073/pnas.2025445118 . ISSN   0027-8424. PMC   8179161 . PMID   34050026.
  9. Chemotaxis Lecture. Uploaded in 2007. available at: http://www.rpgroup.caltech.edu/courses/aph161/2007/lectures/ChemotaxisLecture.pdf Archived 2010-06-19 at the Wayback Machine (Last inspected: 15/04/17)
  10. Élie Metchnikoff". Encyclopædia Britannica. Encyclopædia Britannica, Inc.
  11. Keller-Segel Models for Chemotaxis. 2012. available at: http://www.isn.ucsd.edu/courses/Beng221/problems/2012/BENG221_Project%20-%20Roberts%20Chung%20Yu%20Li.pdf Archived 29 August 2017 at the Wayback Machine (Last inspected by April 2017)
  12. Snyderman R, Gewurz H, Mergenhagen SE (August 1968). "Interactions of the complement system with endotoxic lipopolysaccharide. Generation of a factor chemotactic for polymorphonuclear leukocytes". The Journal of Experimental Medicine. 128 (2): 259–75. doi:10.1084/jem.128.2.259. PMC   2138524 . PMID   4873021.
  13. Adler J, Tso WW (June 1974). ""Decision"-making in bacteria: chemotactic response of Escherichia coli to conflicting stimuli". Science. 184 (4143): 1292–4. Bibcode:1974Sci...184.1292A. doi:10.1126/science.184.4143.1292. PMID   4598187. S2CID   7221477.
  14. Berg, Howard (2004). Berg, Howard C (ed.). E. coli in Motion. Biological and Medical Physics, Biomedical Engineering. Springer. p.  15, 19–29. doi:10.1007/b97370. ISBN   0-387-00888-8. S2CID   35733036.
  15. Yuan J, Fahrner KA, Turner L, Berg HC (July 2010). "Asymmetry in the clockwise and counterclockwise rotation of the bacterial flagellar motor". Proceedings of the National Academy of Sciences of the United States of America. 107 (29): 12846–9. Bibcode:2010PNAS..10712846Y. doi: 10.1073/pnas.1007333107 . PMC   2919929 . PMID   20615986.
  16. "Bacterial Chemotaxis" (PDF). Archived (PDF) from the original on 6 May 2017.
  17. Berg HC, Brown DA (October 1972). "Chemotaxis in Escherichia coli analysed by Three-dimensional Tracking". Nature. 239 (5374): 500–504. Bibcode:1972Natur.239..500B. doi:10.1038/239500a0. PMID   4563019. S2CID   1909173.
  18. Sourjik V, Wingreen NS (April 2012). "Responding to chemical gradients: bacterial chemotaxis". Current Opinion in Cell Biology. 24 (2): 262–268. doi:10.1016/j.ceb.2011.11.008. PMC   3320702 . PMID   22169400.
  19. Berg, Howard C. (1993). Random walks in biology (Expanded, rev. ed.). Princeton, NJ: Princeton Univ. Press. pp. 83–94. ISBN   978-0-691-00064-0.
  20. Sourjik V, Wingreen N (April 2012). "Responding to Chemical Gradients: Bacterial Chemotaxis". Current Opinion in Cell Biology. 24 (2): 262–8. doi:10.1016/j.ceb.2011.11.008. PMC   3320702 . PMID   22169400.
  21. Macnab RM, Koshland DE (September 1972). "The gradient-sensing mechanism in bacterial chemotaxis". Proceedings of the National Academy of Sciences of the United States of America. 69 (9): 2509–12. Bibcode:1972PNAS...69.2509M. doi: 10.1073/pnas.69.9.2509 . PMC   426976 . PMID   4560688.
  22. Nedeljković, Marko; Sastre, Diego; Sundberg, Eric (14 July 2021). "Bacterial Flagellar Filament: A Supramolecular Multifunctional Nanostructure". International Journal of Molecular Sciences. 22 (14): 7521. doi: 10.3390/ijms22147521 . PMC   8306008 . PMID   34299141.
  23. Zhong, Maohua; Yan, Huimin; Li, Yaoming (October 2017). "Flagellin: a unique microbe-associated molecular pattern and a multi-faceted immunomodulator". Cellular & Molecular Immunology. 14 (10): 862–864. doi:10.1038/cmi.2017.78. ISSN   2042-0226. PMC   5649114 . PMID   28845044.
  24. Berg HC (2003). E. coli in motion. New York, NY: Springer. ISBN   978-0-387-00888-2.[ page needed ]
  25. Wadhams, George H.; Armitage, Judith P. (December 2004). "Making sense of it all: bacterial chemotaxis". Nature Reviews Molecular Cell Biology. 5 (12): 1024–1037. doi:10.1038/nrm1524. PMID   15573139. S2CID   205493118.
  26. Galperin, Michael (June 2005). "A census of membrane-bound and intracellular signal transduction proteins in bacteria: Bacterial IQ, extroverts and introverts". BMC Microbiology. 5: 35. doi: 10.1186/1471-2180-5-35 . PMC   1183210 . PMID   15955239.
  27. Gennaro Auletta (2011). Cognitive Biology: Dealing with Information from Bacteria to Minds. United States: Oxford University Press. p. 266. ISBN   978-0-19-960848-5.
  28. 1 2 Falke, Joseph J.; Bass, Randal B.; Butler, Scott L.; Chervitz, Stephen A.; Danielson, Mark A. (1997). "THE TWO-COMPONENT SIGNALING PATHWAY OF BACTERIAL CHEMOTAXIS: A Molecular View of Signal Transduction by Receptors, Kinases, and Adaptation Enzymes". Annual Review of Cell and Developmental Biology. 13: 457–512. doi:10.1146/annurev.cellbio.13.1.457. ISSN   1081-0706. PMC   2899694 . PMID   9442881.
  29. ToxCafe (2 June 2011). "Chemotaxis". Archived from the original on 11 July 2015. Retrieved 23 March 2017 via YouTube.
  30. 1 2 3 4 Wadhams, George H.; Armitage, Judith P. (December 2004). "Making sense of it all: bacterial chemotaxis". Nature Reviews Molecular Cell Biology. 5 (12): 1024–1037. doi:10.1038/nrm1524. ISSN   1471-0080. PMID   15573139. S2CID   205493118.
  31. Shu Chien; Peter C Y Chen; Y C Fung (2008). An Introductory Text to Bioengineering (Advanced Series in Biomechanics - Vol. 4). Singapore: World Scientific Publishing Co. Pte. Ltd. p. 418. ISBN   9789812707932.
  32. Cluzel P, Surette M, Leibler S (March 2000). "An ultrasensitive bacterial motor revealed by monitoring signaling proteins in single cells". Science. 287 (5458): 1652–5. Bibcode:2000Sci...287.1652C. doi:10.1126/science.287.5458.1652. PMID   10698740. S2CID   5334523.
  33. Sourjik V (December 2004). "Receptor clustering and signal processing in E. coli chemotaxis". Trends in Microbiology. 12 (12): 569–76. CiteSeerX   10.1.1.318.4824 . doi:10.1016/j.tim.2004.10.003. PMID   15539117.
  34. Xu F, Bierman R, Healy F, Nguyen H (2016). "A multi-scale model of Escherichia coli chemotaxis from intracellular signaling pathway to motility and nutrient uptake in nutrient gradient and isotropic fluid environments". Computers & Mathematics with Applications. 71 (11): 2466–2478. doi: 10.1016/j.camwa.2015.12.019 .
  35. 1 2 3 4 5 6 Szurmant H, Ordal GW (June 2004). "Diversity in chemotaxis mechanisms among the bacteria and archaea". Microbiology and Molecular Biology Reviews. 68 (2): 301–19. doi:10.1128/MMBR.68.2.301-319.2004. PMC   419924 . PMID   15187186.
  36. 1 2 Yamamoto K, Macnab RM, Imae Y (January 1990). "Repellent response functions of the Trg and Tap chemoreceptors of Escherichia coli". Journal of Bacteriology. 172 (1): 383–8. doi:10.1128/jb.172.1.383-388.1990. PMC   208443 . PMID   2403544.
  37. 1 2 3 Schiffmann E, Corcoran BA, Wahl SM (March 1975). "N-formylmethionyl peptides as chemoattractants for leucocytes". Proceedings of the National Academy of Sciences of the United States of America. 72 (3): 1059–62. Bibcode:1975PNAS...72.1059S. doi: 10.1073/pnas.72.3.1059 . PMC   432465 . PMID   1093163.
  38. Kuruvilla H, Schmidt B, Song S, Bhajjan M, Merical M, Alley C, Griffin C, Yoder D, Hein J, Kohl D, Puffenberger C, Petroff D, Newcomer E, Good K, Heston G, Hurtubise A (2016). "Netrin-1 Peptide Is a Chemorepellent in Tetrahymena thermophila". International Journal of Peptides. 2016: 7142868. doi: 10.1155/2016/7142868 . PMC   4830718 . PMID   27123011.
  39. Hennessey TM (June 2005). "Responses of the ciliates Tetrahymena and Paramecium to external ATP and GTP". Purinergic Signalling. 1 (2): 101–10. doi:10.1007/s11302-005-6213-1. PMC   2096533 . PMID   18404496.
  40. Köhidai L (2016), "Chemotaxis as an Expression of Communication of Tetrahymena", in Witzany G, Nowacki M (eds.), Biocommunication of Ciliates, pp. 65–82, doi:10.1007/978-3-319-32211-7_5, ISBN   978-3-319-32211-7
  41. Berg, H.C.; Purcell, E.M. (November 1977). "Physics of chemoreception". Biophysical Journal. 20 (2): 193–219. Bibcode:1977BpJ....20..193B. doi:10.1016/s0006-3495(77)85544-6. ISSN   0006-3495. PMC   1473391 . PMID   911982.
  42. 1 2 3 4 Levine, Herbert; Rappel, Wouter-Jan (1 February 2013). "The physics of eukaryotic chemotaxis". Physics Today. 66 (2): 24–30. Bibcode:2013PhT....66b..24L. doi:10.1063/PT.3.1884. PMC   3867297 . PMID   24363460.
  43. 1 2 3 4 Vladimirov N, Sourjik V (November 2009). "Chemotaxis: how bacteria use memory". Biological Chemistry. 390 (11): 1097–104. doi:10.1515/BC.2009.130. PMID   19747082. S2CID   207440927.
  44. Ghose, Debraj; Lew, Daniel (1 May 2020). "Mechanistic insights into actin-driven polarity site movement in yeast". Molecular Biology of the Cell. 31 (10): 1085–1102. doi:10.1091/mbc.e20-01-0040. ISSN   1059-1524. PMC   7346724 . PMID   32186970.
  45. 1 2 Skoge M, Yue H, Erickstad M, Bae A, Levine H, Groisman A, Loomis WF, Rappel WJ (October 2014). "Cellular memory in eukaryotic chemotaxis". Proceedings of the National Academy of Sciences of the United States of America. 111 (40): 14448–53. Bibcode:2014PNAS..11114448S. doi: 10.1073/pnas.1412197111 . PMC   4210025 . PMID   25249632.
  46. Kutscher B, Devreotes P, Iglesias PA (February 2004). "Local excitation, global inhibition mechanism for gradient sensing: an interactive applet". Science's STKE. 2004 (219): pl3. doi:10.1126/stke.2192004pl3. PMID   14872096. S2CID   4660870.
  47. Xiong Y, Huang CH, Iglesias PA, Devreotes PN (October 2010). "Cells navigate with a local-excitation, global-inhibition-biased excitable network". Proceedings of the National Academy of Sciences of the United States of America. 107 (40): 17079–86. Bibcode:2010PNAS..10717079X. doi: 10.1073/pnas.1011271107 . PMC   2951443 . PMID   20864631.
  48. Bagorda A, Parent CA (August 2008). "Eukaryotic chemotaxis at a glance". Journal of Cell Science. 121 (Pt 16): 2621–4. CiteSeerX   10.1.1.515.32 . doi:10.1242/jcs.018077. PMC   7213762 . PMID   18685153.
  49. Köhidai L (1999). "Chemotaxis: the proper physiological response to evaluate phylogeny of signal molecules". Acta Biologica Hungarica. 50 (4): 375–94. doi:10.1007/BF03543060. PMID   10735174. S2CID   248703226.
  50. Kedrin D, van Rheenen J, Hernandez L, Condeelis J, Segall JE (September 2007). "Cell motility and cytoskeletal regulation in invasion and metastasis". Journal of Mammary Gland Biology and Neoplasia. 12 (2–3): 143–52. doi:10.1007/s10911-007-9046-4. PMID   17557195. S2CID   31704677.
  51. Solnica-Krezel L, Sepich DS (2012). "Gastrulation: making and shaping germ layers". Annual Review of Cell and Developmental Biology. 28: 687–717. doi:10.1146/annurev-cellbio-092910-154043. PMID   22804578. S2CID   11331182.
  52. Shellard A, Mayor R (July 2016). "Chemotaxis during neural crest migration". Seminars in Cell & Developmental Biology. 55: 111–8. doi:10.1016/j.semcdb.2016.01.031. PMID   26820523.
  53. Becker EL (October 1977). "Stimulated neutrophil locomotion: chemokinesis and chemotaxis". Archives of Pathology & Laboratory Medicine. 101 (10): 509–13. PMID   199132.
  54. Carter SB (January 1967). "Haptotaxis and the mechanism of cell motility". Nature. 213 (5073): 256–60. Bibcode:1967Natur.213..256C. doi:10.1038/213256a0. PMID   6030602. S2CID   4212997.
  55. Kim JY, Haastert PV, Devreotes PN (April 1996). "Social senses: G-protein-coupled receptor signaling pathways in Dictyostelium discoideum". Chemistry & Biology. 3 (4): 239–43. doi: 10.1016/s1074-5521(96)90103-9 . PMID   8807851.
  56. Montell, Craig (November 1999). "Visual Transduction in Drosophila". Annual Review of Cell and Developmental Biology. 15 (1): 231–268. doi:10.1146/annurev.cellbio.15.1.231. PMID   10611962. S2CID   14193715.
  57. Antunes G, Simoes de Souza FM (2016). "Olfactory receptor signaling". G Protein-Coupled Receptors - Signaling, Trafficking and Regulation. Methods in Cell Biology. Vol. 132. pp. 127–45. doi:10.1016/bs.mcb.2015.11.003. ISBN   9780128035955. PMID   26928542.
  58. Thomas, Monica A.; Kleist, Andrew B.; Volkman, Brian F. (2018). "Decoding the chemotactic signal". Journal of Leukocyte Biology. 104 (2): 359–374. doi:10.1002/JLB.1MR0218-044. PMC   6099250 . PMID   29873835.
  59. van Haastert PJ, De Wit RJ, Konijn TM (August 1982). "Antagonists of chemoattractants reveal separate receptors for cAMP, folic acid and pterin in Dictyostelium" (PDF). Experimental Cell Research. 140 (2): 453–6. doi:10.1016/0014-4827(82)90139-2. PMID   7117406. S2CID   27784085.
  60. Witzany, Guenther; Nowacki, Mariusz (2016). Biocommunication of Ciliates. Springer. ISBN   978-3-319-32211-7.[ page needed ]
  61. Köhidai, Laszlo (2016). "Chemotaxis as an Expression of Communication of Tetrahymena". In Witzany, Guenther; Nowacki, Mariusz (eds.). Biocommunication of Ciliates. Springer. pp. 65–82. doi:10.1007/978-3-319-32211-7_5. ISBN   978-3-319-32211-7.
  62. Köhidai L, Csaba G (July 1998). "Chemotaxis and chemotactic selection induced with cytokines (IL-8, RANTES and TNF-alpha) in the unicellular Tetrahymena pyriformis". Cytokine. 10 (7): 481–6. doi:10.1006/cyto.1997.0328. PMID   9702410. S2CID   33755476.
  63. Zigmond SH (November 1977). "Ability of polymorphonuclear leukocytes to orient in gradients of chemotactic factors". The Journal of Cell Biology. 75 (2 Pt 1): 606–16. doi:10.1083/jcb.75.2.606. PMC   2109936 . PMID   264125.
  64. 1 2 3 4 Powell WS, Rokach J (April 2015). "Biosynthesis, biological effects, and receptors of hydroxyeicosatetraenoic acids (HETEs) and oxoeicosatetraenoic acids (oxo-ETEs) derived from arachidonic acid". Biochimica et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 1851 (4): 340–55. doi:10.1016/j.bbalip.2014.10.008. PMC   5710736 . PMID   25449650.
  65. Powell WS, Rokach J (October 2013). "The eosinophil chemoattractant 5-oxo-ETE and the OXE receptor". Progress in Lipid Research. 52 (4): 651–65. doi:10.1016/j.plipres.2013.09.001. PMC   5710732 . PMID   24056189.
  66. Matsuoka T, Narumiya S (September 2007). "Prostaglandin receptor signaling in disease". TheScientificWorldJournal. 7: 1329–47. doi: 10.1100/tsw.2007.182 . PMC   5901339 . PMID   17767353.
  67. Yokomizo T (February 2015). "Two distinct leukotriene B4 receptors, BLT1 and BLT2". Journal of Biochemistry. 157 (2): 65–71. doi:10.1093/jb/mvu078. PMID   25480980.
  68. Sozzani S, Zhou D, Locati M, Bernasconi S, Luini W, Mantovani A, O'Flaherty JT (November 1996). "Stimulating properties of 5-oxo-eicosanoids for human monocytes: synergism with monocyte chemotactic protein-1 and -3". Journal of Immunology. 157 (10): 4664–71. doi: 10.4049/jimmunol.157.10.4664 . PMID   8906847. S2CID   23499393.
  69. Kohidai L, Lang O and Csaba G (2003). "Chemotactic-range-fitting of amino acids and its correlations to physicochemical parameters in Tetrahymena pyriformis - Evolutionary consequences". Cellular and Molecular Biology. 49: OL487–95. PMID   14995080.
  70. 1 2 Murray, James D. (2002). Mathematical Biology I: An Introduction (PDF). Interdisciplinary Applied Mathematics. Vol. 17 (3rd ed.). New York: Springer. pp. 395–417. doi:10.1007/b98868. ISBN   978-0-387-95223-9. Archived (PDF) from the original on 6 May 2022.
  71. Gharasoo M, Centler F, Fetzer I, Thullner M (2014). "How the chemotactic characteristics of bacteria can determine their population patterns". Soil Biology and Biochemistry. 69: 346–358. doi:10.1016/j.soilbio.2013.11.019.
  72. Mackenzie, Dana (6 March 2023). "How animals follow their nose". Knowable Magazine. Annual Reviews. doi: 10.1146/knowable-030623-4 . S2CID   257388244 . Retrieved 13 March 2023.
  73. Reddy, Gautam; Murthy, Venkatesh N.; Vergassola, Massimo (10 March 2022). "Olfactory Sensing and Navigation in Turbulent Environments". Annual Review of Condensed Matter Physics. 13 (1): 191–213. Bibcode:2022ARCMP..13..191R. doi:10.1146/annurev-conmatphys-031720-032754. ISSN   1947-5454. S2CID   243966350.
  74. Lagzi, István (2013). "Chemical Robotics—Chemotactic Drug Carriers". Central European Journal of Medicine. 8 (4): 377–382. doi: 10.2478/s11536-012-0130-9 . S2CID   84150518.
  75. Sengupta S, Dey KK, Muddana HS, Tabouillot T, Ibele ME, Butler PJ, Sen A (January 2013). "Enzyme molecules as nanomotors". Journal of the American Chemical Society. 135 (4): 1406–14. doi:10.1021/ja3091615. PMID   23308365.
  76. Mohajerani, F; Zhao, X; Somasundar, A; Velegol, D; Sen, A (2018). "A Theory of Enzyme Chemotaxis: From Experiments to Modeling". Biochemistry. 57 (43): 6256–6263. arXiv: 1809.02530 . doi:10.1021/acs.biochem.8b00801. PMID   30251529. S2CID   52816076.
  77. Zhao X, Palacci H, Yadav V, Spiering MM, Gilson MK, Butler PJ, Hess H, Benkovic SJ, Sen A (March 2018). "Substrate-driven chemotactic assembly in an enzyme cascade". Nature Chemistry. 10 (3): 311–317. Bibcode:2018NatCh..10..311Z. doi:10.1038/nchem.2905. PMID   29461522.
  78. Guha R, Mohajerani F, Collins M, Ghosh S, Sen A, Velegol D (November 2017). "Chemotaxis of Molecular Dyes in Polymer Gradients in Solution". Journal of the American Chemical Society. 139 (44): 15588–15591. doi:10.1021/jacs.7b08783. PMID   29064685.

Further reading