Ultrasensitivity: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
No edit summary
reformatted refs using citation templates; stripped before running citation bot.
Line 1: Line 1:
In [[molecular biology]], '''ultrasensitivity''' describes an output response that is more sensitive to stimulus change than the hyperbolic [[Michaelis–Menten kinetics| Michaelis-Menten response]]. Ultrasensitivity is one of the [[Biochemical switches in the cell cycle]] and has been implicated in a number of important cellular events, including exiting G2 cell cycle arrests in ''[[Xenopus laevis]]'' oocytes, a stage to which the cell or organism would not want to return.<ref> James E. Ferrell Jr. and Eric M. Machleder, [http://http://www.ncbi.nlm.nih.gov/pubmed/9572732 "The biochemical basis of an all-or-none cell fate switch in Xenopus oocytes."], ''Science'', 1998 May 8;280(5365):895-8 </ref>
In [[molecular biology]], '''ultrasensitivity''' describes an output response that is more sensitive to stimulus change than the hyperbolic [[Michaelis–Menten kinetics| Michaelis-Menten response]]. Ultrasensitivity is one of the [[Biochemical switches in the cell cycle]] and has been implicated in a number of important cellular events, including exiting G2 cell cycle arrests in ''[[Xenopus laevis]]'' oocytes, a stage to which the cell or organism would not want to return.<ref>{{cite journal |pmid=9572732}}</ref>

Ultrasensitivity is a cellular system which triggers entry into a different cellular state<ref>Vivek K Mutalik and KV Venkatesh, "Quantification of the glycogen cascade system: the ultrasensitive responses of liver glycogen synthase and muscle phosphorylase are due to distinctive regulatory designs", Theoretical Biology and Medical Modelling 2005</ref>. Ultrasensitivity gives a small response to first input signal, but an increase in the input signal produces higher and higher levels of output. This acts to filter out noise, as small stimuli and threshold concentrations of the stimulus (input signal) is necessary for the trigger which allows the system to get activated quickly<ref>Eric C. Greenwald, MS and Jeffrey J. Saucerman, PhD,"Bigger, Better, Faster: Principles and Models of AKAP Anchoring Protein Signaling",(J Cardiovasc PharmacolTM 2011;58:462–469)</ref>. Ultrasensitive responses are represented by sigmoidal graphs, which resemble [[cooperativity]]. Quantification of ultrasensitivity is often approximated by the [[Hill equation (biochemistry)]]:
Ultrasensitivity is a cellular system which triggers entry into a different cellular state.<ref>{{cite journal |pmid=15907212}}</ref> Ultrasensitivity gives a small response to first input signal, but an increase in the input signal produces higher and higher levels of output. This acts to filter out noise, as small stimuli and threshold concentrations of the stimulus (input signal) is necessary for the trigger which allows the system to get activated quickly.<ref>{{cite journal |pmid=21562426}}</ref> Ultrasensitive responses are represented by sigmoidal graphs, which resemble [[cooperativity]]. Quantification of ultrasensitivity is often approximated by the [[Hill equation (biochemistry)]]:


'''Response= Stimulus^n/(EC50^n+Stimulus^n)'''
'''Response= Stimulus^n/(EC50^n+Stimulus^n)'''


Where Hill's coefficient (n) may represent quantitative measure of ultrasensitive response<ref name="Ferrell" />.
Where Hill's coefficient (n) may represent quantitative measure of ultrasensitive response.<ref name="Ferrell" />


[[File:Ultrasensitivity.png|thumb|Schematic of an ultrasensitive response (solid line). A Michaelian curve (dashed line) is included for comparison.]]
[[File:Ultrasensitivity.png|thumb|Schematic of an ultrasensitive response (solid line). A Michaelian curve (dashed line) is included for comparison.]]


==Historical development==
==Historical development==
Zero-order ultrasensitivity was first described by Albert Goldbeter and [[Daniel Koshland|Daniel Koshland, Jr]]. in 1981 in a paper in the [[Proceedings of the National Academy of Sciences of the United States of America|Proceedings of the National Academy of Sciences]].<ref name="Goldbeter">{{cite journal|last=Goldbeter|first=Albert|coauthors=Daniel E. Koshland, Jr.|title=An amplified sensitivity arising from covalent modification in biological systems|journal=PNAS|year=1981|volume=78|url=http://www.pnas.org/content/78/11/6840.full.pdf+html}}</ref> They showed using [[Goldbeter-Koshland kinetics|mathematical modeling]] that modification of enzymes operating outside of first order kinetics required only small changes in the concentration of the effector to produce larger changes in the amount of modified protein. This amplification provided added sensitivity in biological control, and implicated the importance of this in many biological systems. Many biological processes are binary (ON-OFF), such as cell fate decisions<ref>Richard I. Joh, Joshua S. Weitz, (March 2011) [http://www.ncbi.nlm.nih.gov/pubmed/21423715 "To Lyse or Not to Lyse: Transient-Mediated Stochastic Fate Determination in Cells Infected by Bacteriophages"], PLoS Comput Biol.</ref>, metabolic states, and signaling pathways. Ultrasensitivity is a switch that helps decision-making in such biological processes.<ref>Anushree Chatterjee, Yiannis N. Kaznessis, (Oct 2008)[http://www.ncbi.nlm.nih.gov/pubmed/18804166 "Tweaking biological switches through a better understanding of bistability behavior"], Curr Opin Biotechnol.</ref>For example, in apoptotic process, a model showed that a positive feedback of inhibition of caspase 3 (Casp3) and Casp9 by inhibitors of apoptosis can bring about ultrasensitivity (bistability). This positive feedback cooperates with Casp3-mediated feedback cleavage of Casp9 to generate irreversibility in caspase activation (switch ON), which leads to cell apoptosis.<ref>Legewie S, Blüthgen N, (Sep 2006) [http://www.ncbi.nlm.nih.gov/pubmed/16978046 "Mathematical modeling identifies inhibitors of apoptosis as mediators of positive feedback and bistability"], PLoS Comput Biol.</ref>Another model also showed similar but different positive feedback controls in Bcl-2 family proteins in apoptotic process.<ref>Jun Cui, Chun Chen, (Jan 2008) [http://www.ncbi.nlm.nih.gov/pubmed/18213378 "Two Independent Positive Feedbacks and Bistability in the Bcl-2 Apoptotic Switch"], PLoS One.</ref>Recently, Jeyeraman et al. have proposed that the phenomenon of ultrasensitivity may be further subdivided into three sub-regimes, separated by sharp stimulus threshold values: OFF, OFF-ON-OFF, and ON. Based on their model, they proposed that this sub-regime of ultrasensitivity, OFF-ON-OFF, is like a switch-like adaption which can be accomplished by coupling N phosphorylation–dephosphorylation cycles unidirectionally, without any explicit feedback loops.<ref name="Jeyaraman">{{cite journal|last=Jeyaraman|first=Srividhya|coauthors=Yongfeng Li, Joseph R. Pomerening|title=Open cascades as simple solutions to providing ultrasensitivity and adaptation in cellular signaling|journal=Phys. Biol.|year=2011|vol=8|url=http://dx.doi.org/10.1088/1478-3975/8/4/046005}}</ref> Other recent work has emphasized that not only is the topology of networks important for creating ultrasensitivite responses, but that their composition (enzymes vs. transcription factors) strongly effects whether they will exhibit robust ultrasensitivity. Mathematical modeling suggests for a broad array of network topologies that a combination of enzymes and transcription factors tends to provide more robust ultransensitivity than that seen in networks composed entirely of transcription factors or composed entirely of enzymes<ref>Shah, N.A., and Sarkar, C.A. [http://www.ploscompbiol.org/article/info%3Adoi%2F10.1371%2Fjournal.pcbi.1002085 Robust network topologies for generating switch-like cellular responses]. PLoS Comput Biol. 2011 Jun;7(6):e1002085</ref>.
Zero-order ultrasensitivity was first described by Albert Goldbeter and [[Daniel Koshland|Daniel Koshland, Jr]]. in 1981 in a paper in the [[Proceedings of the National Academy of Sciences of the United States of America|Proceedings of the National Academy of Sciences]].<ref name="Goldbeter">{{cite journal |first1=Albert |last1=Goldbeter |first2=Daniel E. |last2=Koshland |title=An Amplified Sensitivity Arising from Covalent Modification in Biological Systems |journal=Proceedings of the National Academy of Sciences of the United States of America |jstor=11361 |pmc=349147 |pmid=6947258 |url=http://www.pnas.org/content/78/11/6840.full.pdf+html}}</ref> They showed using [[Goldbeter-Koshland kinetics|mathematical modeling]] that modification of enzymes operating outside of first order kinetics required only small changes in the concentration of the effector to produce larger changes in the amount of modified protein. This amplification provided added sensitivity in biological control, and implicated the importance of this in many biological systems. Many biological processes are binary (ON-OFF), such as cell fate decisions,<ref>{{cite journal |pmid=21423715}}</ref> metabolic states, and signaling pathways. Ultrasensitivity is a switch that helps decision-making in such biological processes.<ref>{{cite journal |pmid=18804166}}</ref> For example, in apoptotic process, a model showed that a positive feedback of inhibition of caspase 3 (Casp3) and Casp9 by inhibitors of apoptosis can bring about ultrasensitivity (bistability). This positive feedback cooperates with Casp3-mediated feedback cleavage of Casp9 to generate irreversibility in caspase activation (switch ON), which leads to cell apoptosis.<ref>{{cite journal |pmid=16978046}}</ref>Another model also showed similar but different positive feedback controls in Bcl-2 family proteins in apoptotic process.<ref>{{cite journal |pmid=18213378}}</ref> Recently, Jeyeraman et al. have proposed that the phenomenon of ultrasensitivity may be further subdivided into three sub-regimes, separated by sharp stimulus threshold values: OFF, OFF-ON-OFF, and ON. Based on their model, they proposed that this sub-regime of ultrasensitivity, OFF-ON-OFF, is like a switch-like adaption which can be accomplished by coupling N phosphorylation–dephosphorylation cycles unidirectionally, without any explicit feedback loops.<ref name="Jeyaraman">{{cite journal |doi=10.1088/1478-3975/8/4/046005}}</ref> Other recent work has emphasized that not only is the topology of networks important for creating ultrasensitivite responses, but that their composition (enzymes vs. transcription factors) strongly effects whether they will exhibit robust ultrasensitivity. Mathematical modeling suggests for a broad array of network topologies that a combination of enzymes and transcription factors tends to provide more robust ultransensitivity than that seen in networks composed entirely of transcription factors or composed entirely of enzymes.<ref>{{cite journal |doi=10.1371/journal.pcbi.1002085}}</ref>


==Development of a Synthetic Ultrasensitive Signaling Pathway==
==Development of a Synthetic Ultrasensitive Signaling Pathway==
Recently it has been shown that a Michaelian signaling pathway can be converted to an ultrasensitive signaling pathway by the introduction of two positive feedback loops <ref>{{cite journal|pmid=21451590}}</ref>. In this synthetic biology approach, Palani and Sarkar began with a linear, graded response pathway, a pathway that showed a proportional increase in signal output relative to the amount of signal input, over a certain range of inputs. This simple pathway was composed of a membrane receptor, a kinase and a transcription factor. Upon activation the membrane receptor phosphorylates the kinase, which moves into the nucleus and phosphorylates the transcription factor, which turns on gene expression. To transform this graded response system into an ultrasensitive, or switch-like signaling pathway, the investigators created two positive feedback loops. In the engineered system, activation of the membrane receptor resulted in increased expression of both the receptor itself and the transcription factor. This was accomplished by placing a promoter specific for this transcription factor upstream of both genes. The authors were able to demonstrate that the synthetic pathway displayed high ultrasensitivity and bistability. They speculate that this type of pathway modification may be generalizable to many similar graded response pathways.
Recently it has been shown that a Michaelian signaling pathway can be converted to an ultrasensitive signaling pathway by the introduction of two positive feedback loops.<ref>{{cite journal |pmid=21451590}}</ref> In this synthetic biology approach, Palani and Sarkar began with a linear, graded response pathway, a pathway that showed a proportional increase in signal output relative to the amount of signal input, over a certain range of inputs. This simple pathway was composed of a membrane receptor, a kinase and a transcription factor. Upon activation the membrane receptor phosphorylates the kinase, which moves into the nucleus and phosphorylates the transcription factor, which turns on gene expression. To transform this graded response system into an ultrasensitive, or switch-like signaling pathway, the investigators created two positive feedback loops. In the engineered system, activation of the membrane receptor resulted in increased expression of both the receptor itself and the transcription factor. This was accomplished by placing a promoter specific for this transcription factor upstream of both genes. The authors were able to demonstrate that the synthetic pathway displayed high ultrasensitivity and bistability. They speculate that this type of pathway modification may be generalizable to many similar graded response pathways.{{fact}}


Recent computational analysis of the effects of a signaling protein's concentration on the presence of an ultrasensitive response has come to complementary conclusions about the influence of a signaling protein's concentration on the conversion of a graded response to an ultrasensitive one. Rather than focus on the generation of signaling proteins through positive feedback, however, the study instead focused on how the dynamics of a signaling protein's exit from the system influences the response. Soyer, Kuwahara, and Csika´sz-Nagy<ref>{{cite journal|pmid=19438711}}</ref> devised a signaling pathway composed of a protein (P) that possesses two possible states (unmodified P or modified P*) and can be modified by an incoming stimulus E. Furthermore, while the unmodified form, P, is permitted to enter or leave the system, P* is only allowed to leave (i.e. it is not generated elsewhere). After varying the parameters of this system, the researchers discovered that the modification of P to P* can shift between a graded response and an ultrasensitive response via the modification of the exit rates of P and P* relative to each other. The transition from an ultrasensitive response to E and a graded response to E was generated when the two rates went from highly similar to highly dissimilar, irrespective of the kinetics of the conversion from P to P* itself. This finding suggests at least two things: 1) the simplifying assumption that the levels of signaling molecules stay constant in a system can severely limit the understanding of ultrasensitivity's complexity; and 2) it may be possible to induce or inhibit ultrasensitivity artificially by regulating the rates of the entry and exit of signaling molecules occupying a system of interest.
Recent computational analysis of the effects of a signaling protein's concentration on the presence of an ultrasensitive response has come to complementary conclusions about the influence of a signaling protein's concentration on the conversion of a graded response to an ultrasensitive one. Rather than focus on the generation of signaling proteins through positive feedback, however, the study instead focused on how the dynamics of a signaling protein's exit from the system influences the response. Soyer, Kuwahara, and Csika´sz-Nagy<ref>{{cite journal |pmid=19438711}}</ref> devised a signaling pathway composed of a protein (P) that possesses two possible states (unmodified P or modified P*) and can be modified by an incoming stimulus E. Furthermore, while the unmodified form, P, is permitted to enter or leave the system, P* is only allowed to leave (i.e. it is not generated elsewhere). After varying the parameters of this system, the researchers discovered that the modification of P to P* can shift between a graded response and an ultrasensitive response via the modification of the exit rates of P and P* relative to each other. The transition from an ultrasensitive response to E and a graded response to E was generated when the two rates went from highly similar to highly dissimilar, irrespective of the kinetics of the conversion from P to P* itself. This finding suggests at least two things: 1) the simplifying assumption that the levels of signaling molecules stay constant in a system can severely limit the understanding of ultrasensitivity's complexity; and 2) it may be possible to induce or inhibit ultrasensitivity artificially by regulating the rates of the entry and exit of signaling molecules occupying a system of interest.


==Mechanisms==
==Mechanisms==
Ultrasensitivity can be achieved through several mechanisms:
Ultrasensitivity can be achieved through several mechanisms:
#Multistep mechanisms (examples: cooperativity)<ref>{{cite journal |pmid=12023217}}</ref> and multisite phosphorylation<ref>{{cite journal |pmid=14744999}}</ref>
#Multistep mechanisms (examples: cooperativity)<ref>{{cite journal|last=Thattai|first=Mukund|coauthors=Alexandar van Oudenaardan|title=Attenuation of Noise in Ultrasensitive Signaling Cascades|journal=Biophysical Journal|year=2002|month=June|volume=82|pages=2943–295|url=http://www.ncbi.nlm.nih.gov/pmc/articles/PMC1302082/pdf/12023217.pdf}}</ref> and multisite phosphorylation<ref>{{cite journal|last=Markevich|first=Nick|coauthors=Jan Hoek, and Boris N. Kholodenko|title=Signaling switches and bistability arising from multisite phosphorylation in protein kinase cascades|journal=Journal of Cell Biology|year=2004|month=February|volume=164|issue=3|pages=353-359|url=http://jcb.rupress.org/content/164/3/353.abstract}}</ref>
#Buffering mechanisms (examples: decoy phosphorylation sites)<ref>{{cite journal|last=Smith|first=Nicholas R.|coauthors=Kenneth E. Prehoda|title=Robust Spindle Alignment in Drosophila Neuroblasts by Ultrasensitive Activation of Pins|journal=Molecular Cell|year=2011|month=August|volume=43|issue=4|pages=540-549|doi=doi:10.1016/j.molcel.2011.06.030|url=http://www.sciencedirect.com/science/article/pii/S1097276511005776}}</ref> or stoichiometric inhibitors<ref>{{cite journal|last=Kim|first=Sun Young|coauthors=James E. Ferrell, Jr.|title=Substrate Competition as a Source of Ultrasensitivity in the Inactivation of Wee1|journal=Cell|year=2007|month=March|volume=128|issue=6|pages=1133-1145|doi=doi:10.1016/j.cell.2007.01.039|url=http://www.sciencedirect.com/science/article/pii/S0092867407002024}}</ref>
#Buffering mechanisms (examples: decoy phosphorylation sites)<ref>{{cite journal |pages=540-9 |doi=10.1016/j.molcel.2011.06.030}}</ref> or stoichiometric inhibitors<ref>{{cite journal |pages=1133-45 |doi=10.1016/j.cell.2007.01.039}}</ref>
#Changes in localisation (such as translocation across the nuclear envelope)
#Changes in localisation (such as translocation across the nuclear envelope)
#Saturation mechanisms (also known as zero-order ultrasensitivity)<ref>{{cite journal|last=Huang|first=Chi-Ying F.|coauthors=James E. Ferrell, Jr.|title=Ultrasensitivity in the mitogen-activated protein kinase cascade|journal=PNAS|year=1996|month=September|volume=93|pages=10078-10083|url=http://www.pnas.org/content/93/19/10078.full.pdf}}</ref>
#Saturation mechanisms (also known as zero-order ultrasensitivity)<ref>{{cite journal |pmi8816754}}</ref>
#Positive feedback<ref name="Sneppen">{{cite journal|last=Sneppen|first=Kim|coauthors=Mille A Micheelsen and Ian B Dodd|title=Ultrasensitive gene regulation by positive feedback loops in nucleosome modification|journal=Mol Syst Biol|year=2008|volume=4|pages=182|doi=10.1038/msb.2008.21|url=http://www.ncbi.nlm.nih.gov/pmc/articles/PMC2387233/}}</ref>
#Positive feedback<ref name="Sneppen">{{cite journal |doi=10.1038/msb.2008.21}}</ref>
#Allovalency
#Allovalency
#Non-Zero-Order Ultrasensitivity in Membrane Proteins
#Non-Zero-Order Ultrasensitivity in Membrane Proteins
Line 28: Line 29:


===Multistep Mechanisms===
===Multistep Mechanisms===
Multipstep ultrasensitivity occurs when a single effector acts on several steps in a cascade<ref>Albert Goldbeter and Daniel E. Koshland, Jr Ultrasensitivity in Biochemical Systems Controlled by Covalent Modification</ref>. Successive cascade signals can result in higher levels of noise being introduced into the signal that can interfere with the final output. This is especially relevant for large cascades, such as the flagellar regulatory system in which the master regulator signal is transmitted through multiple intermediate regulators before activating transcription (<ref>Kalir S, McClure J, Pabbaraju K, Southward C, Ronen M, Leibler S, Surette MG, Alon U. Ordering genes in a flagella pathway by analysis of expression kinetics from living bacteria. Science. 2001 Jun 15;292(5524):2080-3. PubMed PMID: 11408658.</ref>). Cascade ultrasensitivity can reduce noise and therefore require less input for activation<ref>{{cite journal|last=Thattai|first=Mukund|coauthors=Alexandar van Oudenaardan|title=Attenuation of Noise in Ultrasensitive Signaling Cascades|journal=Biophysical Journal|year=2002|month=June|volume=82|pages=2943–295|url=http://www.ncbi.nlm.nih.gov/pmc/articles/PMC1302082/pdf/12023217.pdf}}</ref>. Additionally, multiple phosphorylation events are an example of ultrasensitivity. Recent modeling has shown that multiple phosphorylation sites on membrane proteins could serve to locally saturate enzyme activity. Proteins at the membrane are greatly reduced in mobility compared to those in the cytoplasm, this means that a membrane tethered enzyme acting upon a membrane protein will take longer to diffuse away. With the addition of multiple phosphorylation sites upon the membrane substrate, the enzyme can - by a combination of increased local concentration of enzyme and increased substrates - quickly reach saturation <ref>Omer Dushek, P. Anton van der Merwe, Vahid Shahrezael Ultrasensitivity in Multisite Phosphorylation of Membrane Anchored Proteins</ref>.
Multipstep ultrasensitivity occurs when a single effector acts on several steps in a cascade.<ref>{{cite journal |pmid=6501300}}</ref> Successive cascade signals can result in higher levels of noise being introduced into the signal that can interfere with the final output. This is especially relevant for large cascades, such as the flagellar regulatory system in which the master regulator signal is transmitted through multiple intermediate regulators before activating transcription.<ref>{{cite journal |pmid=11408658}}</ref> Cascade ultrasensitivity can reduce noise and therefore require less input for activation.<ref>{{cite journal |pmid=12023217}}</ref> Additionally, multiple phosphorylation events are an example of ultrasensitivity. Recent modeling has shown that multiple phosphorylation sites on membrane proteins could serve to locally saturate enzyme activity. Proteins at the membrane are greatly reduced in mobility compared to those in the cytoplasm, this means that a membrane tethered enzyme acting upon a membrane protein will take longer to diffuse away. With the addition of multiple phosphorylation sites upon the membrane substrate, the enzyme can - by a combination of increased local concentration of enzyme and increased substrates - quickly reach saturation.<ref>{{cite journal |pmid=21354391}}</ref>


===Buffering Mechanisms===
===Buffering Mechanisms===
Buffering Mechanisms such as molecular [[Titration|titration]] can generate ultrasensitivity. ''[[In vitro]]'', this can be observed for the simple mechanism:
Buffering Mechanisms such as molecular [[Titration|titration]] can generate ultrasensitivity. ''[[In vitro]]'', this can be observed for the simple mechanism:
:<math>\ A +\ B \rightleftharpoons \ AB</math>
:<math>\ A +\ B \rightleftharpoons \ AB</math>
Where the monomeric form of A is active and it can be inactivated by binding B to form the heterodimer AB. When the concentration of <math>B_{T}</math> (<math>B_{T}</math> = [B] + [AB]) is much greater than the [[Dissociation constant|<math>K_{d}</math>]], this system exhibits a threshold determined by the concentration of <math>B_{T}</math>.<ref>{{cite journal|last=McCarrey|first=J. R.|coauthors=Riggs, A.D.|title=Determinator-inhibitor pairs as a mechanism for threshold setting in development: a possible function for pseudogenes|journal=Proc Natl Acad Sci|year=1986|month=February|volume=83|issue=3|pages=679-683|pmid=2418440}}</ref> At concentrations of <math>A_{T}</math> (<math>A_{T}</math> = [A] +[AB]), lower than <math>B_{T}</math>, B acts as a buffer to free A and nearly all A will be found as AB. However, at the equivalence point, when <math>A_{T}</math> ≈ <math>B_{T}</math>, <math>B_{T}</math> can no longer buffer the increase in <math>A_{T}</math>, so a small increase in <math>A_{T}</math> causes a large increase in A.<ref name="Buchler 2008">{{cite journal|last=Buchler|first=N. E.|coauthors=Louis, M.|title=Molecular Titration and Ultrasensitivity in Regulatory Networks|journal=J Mol Biol|year=2008|volume=384|issue=5|pages=1106-1119|pmid=18938177}}</ref> The strength of the ultrasensitivity of [A] to changes in <math>A_{T}</math> is determined by <math>B_{T}</math>/<math>K_{d}</math>.<ref name="Buchler 2008" /> Ultrasensitivity occurs when this ratio is greater than one and is increased as the ratio increases. Above the equivalence point, <math>A_{T}</math> and A are again linearly related.
Where the monomeric form of A is active and it can be inactivated by binding B to form the heterodimer AB. When the concentration of <math>B_{T}</math> (<math>B_{T}</math> = [B] + [AB]) is much greater than the [[Dissociation constant|<math>K_{d}</math>]], this system exhibits a threshold determined by the concentration of <math>B_{T}</math>.<ref>{{cite journal |pmid=2418440}}</ref> At concentrations of <math>A_{T}</math> (<math>A_{T}</math> = [A] +[AB]), lower than <math>B_{T}</math>, B acts as a buffer to free A and nearly all A will be found as AB. However, at the equivalence point, when <math>A_{T}</math> ≈ <math>B_{T}</math>, <math>B_{T}</math> can no longer buffer the increase in <math>A_{T}</math>, so a small increase in <math>A_{T}</math> causes a large increase in A.<ref name="Buchler 2008">{{cite journal |pmid=18938177}}</ref> The strength of the ultrasensitivity of [A] to changes in <math>A_{T}</math> is determined by <math>B_{T}</math>/<math>K_{d}</math>.<ref name="Buchler 2008" /> Ultrasensitivity occurs when this ratio is greater than one and is increased as the ratio increases. Above the equivalence point, <math>A_{T}</math> and A are again linearly related.
''[[In vivo]]'', the synthesis of A and B as well as the degradation of all three components complicates generation of ultrasensitivity. If the synthesis rates of A and B are equal this system still exhibits ultrasensitivity at the equivalence point.<ref name="Buchler 2008" />
''[[In vivo]]'', the synthesis of A and B as well as the degradation of all three components complicates generation of ultrasensitivity. If the synthesis rates of A and B are equal this system still exhibits ultrasensitivity at the equivalence point.<ref name="Buchler 2008" />


One example of a buffering mechanism is protein sequestration, which is a common mechanism found in signalling and regulatory networks.<ref> Buchler, NE, and Cross FR. 2009. Protein sequestration generates a flexible ultrasensitive response in a genetic network. "Molecular Systems Biology" 5 (272) </ref> In 2009, Buchler and Cross constructed a synthetic genetic network that was regulated by protein sequestration of a transcriptional activator by a dominant-negative inhibitor. They showed that this system results in a flexibile ultrasensitive response in gene expression. It is flexible in that the degree of ultrasensitivity can be altered by changing expression levels of the dominant-negative inhibitor. Figure 1 in their article illustrates how an active transcription factor can be sequestered by an inhibitor into the inactive complex AB that is unable to bind DNA. This type of mechanism results in an “all-or-none” response, or ultransensitivy, when the concentration of the regulatory protein increases to the point of depleting the inhibitor. Robust buffering against a response exists below this concentration threshold , and when it is reached any small increase in input is amplified into a large change in output.
One example of a buffering mechanism is protein sequestration, which is a common mechanism found in signalling and regulatory networks.<ref>{{cite journal |pmid=19455136}}</ref> In 2009, Buchler and Cross constructed a synthetic genetic network that was regulated by protein sequestration of a transcriptional activator by a dominant-negative inhibitor. They showed that this system results in a flexibile ultrasensitive response in gene expression. It is flexible in that the degree of ultrasensitivity can be altered by changing expression levels of the dominant-negative inhibitor. Figure 1 in their article illustrates how an active transcription factor can be sequestered by an inhibitor into the inactive complex AB that is unable to bind DNA. This type of mechanism results in an “all-or-none” response, or ultransensitivy, when the concentration of the regulatory protein increases to the point of depleting the inhibitor. Robust buffering against a response exists below this concentration threshold, and when it is reached any small increase in input is amplified into a large change in output.{{fact}}


=== Changes in localization ===
=== Changes in localization ===


'''Translocation'''
====Translocation====

Signal transduction is regulated in various ways and one of the way is translocation. Regulated translocation generates ultrasensitive response in mainly three ways,
1) Regulated translocation increases the local concentration of the signaling protein. When concentration of the signaling protein is high enough to partially saturate the enzyme that inactivates it, ultrasensitive response is generated.
2) Translocation of multiple components of the signaling cascade where stimulus (input signal) causes translocation of both, signaling protein and its activator in the same subcellular compartment and thereby generates ultrasensitive response which increases speed and accuracy of the signal.
3) Translocation to the compartment which contains stoichiometric inhibitors.<ref name="Ferrell">James E. Ferrell Jr., "How regulated protein translocation can produce switch-like responses", TIBS 23 – DECEMBER 1998</ref>



Signal transduction is regulated in various ways and one of the way is translocation. Regulated translocation generates ultrasensitive response in mainly three ways:
#Regulated translocation increases the local concentration of the signaling protein. When concentration of the signaling protein is high enough to partially saturate the enzyme that inactivates it, ultrasensitive response is generated.
#Translocation of multiple components of the signaling cascade where stimulus (input signal) causes translocation of both, signaling protein and its activator in the same subcellular compartment and thereby generates ultrasensitive response which increases speed and accuracy of the signal.
#Translocation to the compartment which contains stoichiometric inhibitors.<ref name="Ferrell">{{cite journal |pmid=9868363}}</ref>


Translocation is one of way regulating signal transduction and it generates ultrasensitive response such as switch-like response and multistep-feedback loop mechanism.
Translocation is one of way regulating signal transduction and it generates ultrasensitive response such as switch-like response and multistep-feedback loop mechanism.
A switch-like response will occur if translocation raises the local concentration of a signaling protein.
A switch-like response will occur if translocation raises the local concentration of a signaling protein.
For example, [[Epidermal Growth Factor]] (EGF) receptor can be internalized through clathrin-independent endocytosis (CIE) and/or clathrin-independent endoxytosis (CDE) in ligand concentration dependent manner. the distribution of receptors into the two pathways was shown to be EGF-concentration dependent. In the presence of low concentrations of EGF, the receptor was exclusively internalized via CDE, whereas at high oncentrations, receptors were equally distributed between CDE and CIE (Fig.1).
For example, [[Epidermal Growth Factor]] (EGF) receptor can be internalized through clathrin-independent endocytosis (CIE) and/or clathrin-independent endoxytosis (CDE) in ligand concentration dependent manner. the distribution of receptors into the two pathways was shown to be EGF-concentration dependent. In the presence of low concentrations of EGF, the receptor was exclusively internalized via CDE, whereas at high oncentrations, receptors were equally distributed between CDE and CIE (Fig.1).<ref>{{cite journal |pmid=18394191}}</ref><ref>{{cite journal |pmid=9868363}}</ref>
<ref>Hannah Schmidt-Glenewinkel, Ivayla Vacheva, Daniela Hoeller, Ivan Dikic and Roland Eils, “An ultrasensitive sorting mechanism for EGF Receptor Endocytosis”, BMC Systems Biology 2008.</ref><ref>James E. Ferrell Jr., "How regulated protein translocation can produce switch-like responses", TiBS 1998</ref>

<gallery></gallery>


===Saturation mechanisms (Zero-order ultrasensitivity)===
===Saturation mechanisms (Zero-order ultrasensitivity)===

Revision as of 13:56, 5 December 2011

In molecular biology, ultrasensitivity describes an output response that is more sensitive to stimulus change than the hyperbolic Michaelis-Menten response. Ultrasensitivity is one of the Biochemical switches in the cell cycle and has been implicated in a number of important cellular events, including exiting G2 cell cycle arrests in Xenopus laevis oocytes, a stage to which the cell or organism would not want to return.[1]

Ultrasensitivity is a cellular system which triggers entry into a different cellular state.[2] Ultrasensitivity gives a small response to first input signal, but an increase in the input signal produces higher and higher levels of output. This acts to filter out noise, as small stimuli and threshold concentrations of the stimulus (input signal) is necessary for the trigger which allows the system to get activated quickly.[3] Ultrasensitive responses are represented by sigmoidal graphs, which resemble cooperativity. Quantification of ultrasensitivity is often approximated by the Hill equation (biochemistry):

Response= Stimulus^n/(EC50^n+Stimulus^n)

Where Hill's coefficient (n) may represent quantitative measure of ultrasensitive response.[4]

Schematic of an ultrasensitive response (solid line). A Michaelian curve (dashed line) is included for comparison.

Historical development

Zero-order ultrasensitivity was first described by Albert Goldbeter and Daniel Koshland, Jr. in 1981 in a paper in the Proceedings of the National Academy of Sciences.[5] They showed using mathematical modeling that modification of enzymes operating outside of first order kinetics required only small changes in the concentration of the effector to produce larger changes in the amount of modified protein. This amplification provided added sensitivity in biological control, and implicated the importance of this in many biological systems. Many biological processes are binary (ON-OFF), such as cell fate decisions,[6] metabolic states, and signaling pathways. Ultrasensitivity is a switch that helps decision-making in such biological processes.[7] For example, in apoptotic process, a model showed that a positive feedback of inhibition of caspase 3 (Casp3) and Casp9 by inhibitors of apoptosis can bring about ultrasensitivity (bistability). This positive feedback cooperates with Casp3-mediated feedback cleavage of Casp9 to generate irreversibility in caspase activation (switch ON), which leads to cell apoptosis.[8]Another model also showed similar but different positive feedback controls in Bcl-2 family proteins in apoptotic process.[9] Recently, Jeyeraman et al. have proposed that the phenomenon of ultrasensitivity may be further subdivided into three sub-regimes, separated by sharp stimulus threshold values: OFF, OFF-ON-OFF, and ON. Based on their model, they proposed that this sub-regime of ultrasensitivity, OFF-ON-OFF, is like a switch-like adaption which can be accomplished by coupling N phosphorylation–dephosphorylation cycles unidirectionally, without any explicit feedback loops.[10] Other recent work has emphasized that not only is the topology of networks important for creating ultrasensitivite responses, but that their composition (enzymes vs. transcription factors) strongly effects whether they will exhibit robust ultrasensitivity. Mathematical modeling suggests for a broad array of network topologies that a combination of enzymes and transcription factors tends to provide more robust ultransensitivity than that seen in networks composed entirely of transcription factors or composed entirely of enzymes.[11]

Development of a Synthetic Ultrasensitive Signaling Pathway

Recently it has been shown that a Michaelian signaling pathway can be converted to an ultrasensitive signaling pathway by the introduction of two positive feedback loops.[12] In this synthetic biology approach, Palani and Sarkar began with a linear, graded response pathway, a pathway that showed a proportional increase in signal output relative to the amount of signal input, over a certain range of inputs. This simple pathway was composed of a membrane receptor, a kinase and a transcription factor. Upon activation the membrane receptor phosphorylates the kinase, which moves into the nucleus and phosphorylates the transcription factor, which turns on gene expression. To transform this graded response system into an ultrasensitive, or switch-like signaling pathway, the investigators created two positive feedback loops. In the engineered system, activation of the membrane receptor resulted in increased expression of both the receptor itself and the transcription factor. This was accomplished by placing a promoter specific for this transcription factor upstream of both genes. The authors were able to demonstrate that the synthetic pathway displayed high ultrasensitivity and bistability. They speculate that this type of pathway modification may be generalizable to many similar graded response pathways.[citation needed]

Recent computational analysis of the effects of a signaling protein's concentration on the presence of an ultrasensitive response has come to complementary conclusions about the influence of a signaling protein's concentration on the conversion of a graded response to an ultrasensitive one. Rather than focus on the generation of signaling proteins through positive feedback, however, the study instead focused on how the dynamics of a signaling protein's exit from the system influences the response. Soyer, Kuwahara, and Csika´sz-Nagy[13] devised a signaling pathway composed of a protein (P) that possesses two possible states (unmodified P or modified P*) and can be modified by an incoming stimulus E. Furthermore, while the unmodified form, P, is permitted to enter or leave the system, P* is only allowed to leave (i.e. it is not generated elsewhere). After varying the parameters of this system, the researchers discovered that the modification of P to P* can shift between a graded response and an ultrasensitive response via the modification of the exit rates of P and P* relative to each other. The transition from an ultrasensitive response to E and a graded response to E was generated when the two rates went from highly similar to highly dissimilar, irrespective of the kinetics of the conversion from P to P* itself. This finding suggests at least two things: 1) the simplifying assumption that the levels of signaling molecules stay constant in a system can severely limit the understanding of ultrasensitivity's complexity; and 2) it may be possible to induce or inhibit ultrasensitivity artificially by regulating the rates of the entry and exit of signaling molecules occupying a system of interest.

Mechanisms

Ultrasensitivity can be achieved through several mechanisms:

  1. Multistep mechanisms (examples: cooperativity)[14] and multisite phosphorylation[15]
  2. Buffering mechanisms (examples: decoy phosphorylation sites)[16] or stoichiometric inhibitors[17]
  3. Changes in localisation (such as translocation across the nuclear envelope)
  4. Saturation mechanisms (also known as zero-order ultrasensitivity)[18]
  5. Positive feedback[19]
  6. Allovalency
  7. Non-Zero-Order Ultrasensitivity in Membrane Proteins
  8. Dissipative Allostery

Multistep Mechanisms

Multipstep ultrasensitivity occurs when a single effector acts on several steps in a cascade.[20] Successive cascade signals can result in higher levels of noise being introduced into the signal that can interfere with the final output. This is especially relevant for large cascades, such as the flagellar regulatory system in which the master regulator signal is transmitted through multiple intermediate regulators before activating transcription.[21] Cascade ultrasensitivity can reduce noise and therefore require less input for activation.[22] Additionally, multiple phosphorylation events are an example of ultrasensitivity. Recent modeling has shown that multiple phosphorylation sites on membrane proteins could serve to locally saturate enzyme activity. Proteins at the membrane are greatly reduced in mobility compared to those in the cytoplasm, this means that a membrane tethered enzyme acting upon a membrane protein will take longer to diffuse away. With the addition of multiple phosphorylation sites upon the membrane substrate, the enzyme can - by a combination of increased local concentration of enzyme and increased substrates - quickly reach saturation.[23]

Buffering Mechanisms

Buffering Mechanisms such as molecular titration can generate ultrasensitivity. In vitro, this can be observed for the simple mechanism:

Where the monomeric form of A is active and it can be inactivated by binding B to form the heterodimer AB. When the concentration of ( = [B] + [AB]) is much greater than the , this system exhibits a threshold determined by the concentration of .[24] At concentrations of ( = [A] +[AB]), lower than , B acts as a buffer to free A and nearly all A will be found as AB. However, at the equivalence point, when , can no longer buffer the increase in , so a small increase in causes a large increase in A.[25] The strength of the ultrasensitivity of [A] to changes in is determined by /.[25] Ultrasensitivity occurs when this ratio is greater than one and is increased as the ratio increases. Above the equivalence point, and A are again linearly related. In vivo, the synthesis of A and B as well as the degradation of all three components complicates generation of ultrasensitivity. If the synthesis rates of A and B are equal this system still exhibits ultrasensitivity at the equivalence point.[25]

One example of a buffering mechanism is protein sequestration, which is a common mechanism found in signalling and regulatory networks.[26] In 2009, Buchler and Cross constructed a synthetic genetic network that was regulated by protein sequestration of a transcriptional activator by a dominant-negative inhibitor. They showed that this system results in a flexibile ultrasensitive response in gene expression. It is flexible in that the degree of ultrasensitivity can be altered by changing expression levels of the dominant-negative inhibitor. Figure 1 in their article illustrates how an active transcription factor can be sequestered by an inhibitor into the inactive complex AB that is unable to bind DNA. This type of mechanism results in an “all-or-none” response, or ultransensitivy, when the concentration of the regulatory protein increases to the point of depleting the inhibitor. Robust buffering against a response exists below this concentration threshold, and when it is reached any small increase in input is amplified into a large change in output.[citation needed]

Changes in localization

Translocation

Signal transduction is regulated in various ways and one of the way is translocation. Regulated translocation generates ultrasensitive response in mainly three ways:

  1. Regulated translocation increases the local concentration of the signaling protein. When concentration of the signaling protein is high enough to partially saturate the enzyme that inactivates it, ultrasensitive response is generated.
  2. Translocation of multiple components of the signaling cascade where stimulus (input signal) causes translocation of both, signaling protein and its activator in the same subcellular compartment and thereby generates ultrasensitive response which increases speed and accuracy of the signal.
  3. Translocation to the compartment which contains stoichiometric inhibitors.[4]

Translocation is one of way regulating signal transduction and it generates ultrasensitive response such as switch-like response and multistep-feedback loop mechanism. A switch-like response will occur if translocation raises the local concentration of a signaling protein. For example, Epidermal Growth Factor (EGF) receptor can be internalized through clathrin-independent endocytosis (CIE) and/or clathrin-independent endoxytosis (CDE) in ligand concentration dependent manner. the distribution of receptors into the two pathways was shown to be EGF-concentration dependent. In the presence of low concentrations of EGF, the receptor was exclusively internalized via CDE, whereas at high oncentrations, receptors were equally distributed between CDE and CIE (Fig.1).[27][28]

Saturation mechanisms (Zero-order ultrasensitivity)

Zero-order ultrasensitivity takes place under saturating conditions.[29] For example, consider an enzymatic step with a kinase, phosphatase, and substrate. Steady state levels of the phosphorylated substrate have an ultrasensitive response when there is enough substrate to saturate all available kinases and phosphatases.[29][30] Under these conditions, small changes in the ratio of kinase to phosphatase activity can dramatically change the number of phosphorylated substrate (For a graph illustrating this behavior, see [5]). This enhancement in sensitivity of steady state phosphorylated substrate to Km, or the ratio of kinase to phosphatase activity, is termed zero-order to distinguish it from the first order behavior described by Michaelis-Menten dynamics, wherein the steady state concentration responds in a more gradual fashion than the switch-like behavior exhibited in ultrasensitivity.[31]

Using the notation from Goldbeter & Koshland[5] , let W be a certain substrate protein and let W' be a covalently modified version of W. The conversion of W to W' is catalyzed by some enzyme E1 and the reverse conversion of W' to W is catalyzed by a second enzyme E2 according to following equations:


[1] W + E1 WE1 W' + E1


[2] W' + E2 W'E2 W + E2


The concentrations of all necessary components (such as ATP) are assumed to be constant and represented in the kinetic constants. Using equations 1 and 2, the kinetic equations of appearance over time for each component are:


[3] = -a1[W][E1] + d1[WE1] + k2[W'E2]


[4] = a1[W][E1] - (d1+k1)[WE1]


[5] = -a2[W'][E2] + d2[W'E2] + k1[WE1]


[6] = a2[W'][E2] - (d2+k2)[W'E2]


The the total concentration of each component is given by:


[7] Wt = [W] + [W'] + [WE1] +[W'E2]


[8] E1t = [E1] + [WE1]


[9] E2t = [E2] + [W'E2]


The zero order mechanism assumes that the [Wt] >> [E1] or [E2]. In other words the system is in a Michaelis-Menten steady state, which means, to a good approximation, [WE1] and [W'E2] are constant. From these kinetic expressions one can solve for V1/V2 at steady state where k1[WE1]=k2[W'E2] and W = 1 – W'.


[10]


Where:


[11]


[12]


[13]


[14]


When the V1/V2 is plotted against the molar ratio of W' (W'/Wt) and W (W/Wt) it can be seen that the W to W' conversion occurs over a much smaller change in the V1/V2 ratio than it would under first order (non-saturating) conditions, which is the telltale sign of ultrasensitivity.

Positive Feedback

Positive feedback loops can cause ultrasensitive responses. An example of this is seen in the transcription of certain eukaryotic genes in which non-cooperative transcription factor binding changes positive feedback loops of histone modification that results in an ultrasensitive activation of transcription. The binding of a transcription factor recruits histone acetyltransferases and methyltransferases. The acetylation and methylation of histones recruits more acetyltransferases and methyltransferases that results in a positive feedback loop. Ultimately, this results in activation of transcription.[19]

Additionally, positive feedback can induce bistability in Cyclin B1- by the two regulators Wee1 and Cdc25C, leading to the cell's decision to commit to mitosis. The system cannot be stable at intermediate levels of Cyclin B1, and the transition between the two stable states is abrupt when increasing levels of Cyclin B1 switches the system from low to high activity. Exhibiting hysteresis, for different levels of Cyclin B1, the switches from low to high and high to low states vary[32] . However, the emergence of a bistable system is highly influenced by the sensitivity of its feedback loops. It has been shown in "Xenopus" egg extracts that Cdc25C hyperphosphorylation is a highly ultrasensitive function of Cdk activity, displaying a high value of the Hill coefficient (approx. 11), and the dephosphorylation step of Ser 287 in Cdc25C (also involved in Cdc25C activation) is even more ultrasensitive, displaying a Hill coefficient of approx. 32[33] .

File:Cyclin.jpg
Model of the regulation of activation of the Cyclin B-Cdk complex involving reciprocal positive (Cdc25C) and double-negative (Wee1) feedback loops.

[32]

Allovalency

A proposed mechanism of ultrasensitvity, called allovalency, suggests that activity "derives from a high local concentration of interaction sites moving independently of each other"[34] Allovalency was fist proposed when it was believed to occur in the pathway in which Sic1, is degraded in order for Cdk1-Clb (B-type cyclins) to allow entry into mitosis. Sic1 must be phosphorylated multiple times in order to be recognized and degraded by Cdc4 of the SCF Complex.[35] Since Cdc4 only has one recognition site for these phosphorylated residues it was suggested that as the amount of phosphorylation increases, it exponentially increases the likelihood that Sic1 is recognized and degraded by Cdc4. This type of interaction was thought to be relatively immune to loss of any one site and easily tuned to any given threshold by adjusting the properties of individual sites. Assumptions for the allovalency mechanism were based off of a general mathematical model that describes the interaction between a polyvalent disordered ligand and a single receptor site[36] File:Principlesofallovalentinteractions.jpeg It was later found that the ultrasentivity in Cdk1 levels by degregation of Sic1 is in fact due to a positive feedback loop.[37]

Non-Zero-Order Ultrasensitivity in Membrane Proteins

Modeling by Dushek et. al.[38] proposes a possible mechanism for ultrasensitivity outside of the zero-order regime. For the case of membrane-bound enzymes acting on membrane-bound substrates with multiple enzymatic sites (such as tyrosine-phosphorylated receptors like the T-Cell receptor), ultrasensitive responses could be seen, crucially dependent on three factors: 1) limited diffusion in the membrane, 2) multiple binding sites on the substrate, and 3) brief enzymatic inactivation following catalysis.

Under these particular conditions, although the enzyme may be in excess of the substrate (first-order regime), the enzyme is effectively locally saturated with substrate due to the multiple binding sites, leading to switch-like responses. This mechanism of ultrasensitivity is independent of enzyme concentration, however the signal is significantly enhanced depending on the number of binding sites on the substrate.[39] Both conditional factors (limited diffusion and inactivation) are physiologically plausible, but have yet to be experimentally confirmed. Dushek’s modeling found increasing Hill cooperativity numbers with more substrate sites (phosphorylation sites), and with greater steric/diffusional hindrance between enzyme and substrate. This mechanism of ultrasensitivity based on local enzyme saturation arises partly from passive properties of slow membrane diffusion, and therefore may be generally applicable.

Dissipative Allostery

The bacterial flagellar motor has been proposed to follow a dissipative allosteric model, where ultrasensitivity comes as a combination of protein binding affinity and energy contributions from the proton motive force (see Flagellar motors and chemotaxis below).

Hill Coefficient

Ultrasensitive behavior is typically represented by a sigmoidal curve, as small alterations in the stimulus can trigger large changes in the response. The Hill coefficient quantifies the steepness of a sigmoidal stimulus-response curve and it is therefore a sensitivity parameter. It is often used to assess the cooperativity of a system. A Hill coefficient greater than one is indicative of positive cooperativity and thus, the system exhibits ultrasensitivity. [40] Systems with a Hill coefficient of 1 are noncooperative and follow the classical Michaelis-Menten kinetics. Enzymes exhibiting noncooperative activity are represented by hyperbolic stimulus/response curves, compared to sigmoidal curves for cooperative (ultrasensitive) enzymes.[41]

Hill equation:

θ = =
Where n = Hill Coefficient

In mitogen-activated protein kinase (MAPK) signaling (see example below), the ultrasensitivity of the signaling is supported by the sigmoidal stimulus/response curve that is comparable to an enzyme with a Hill coefficient of 4.0-5.0. This is even more ultrasensitive to the cooperative binding activity of hemoglobin, which has a Hill coefficient of 2.8.[41]

Role in Cellular Processes

MAP Kinase Signaling Cascade

A ubiquitous signaling motif that exhibits ultrasensitivity is the MAPK (mitogen-activated protein kinase) cascade, which can take a graded input signal and produce a switch-like output, such as gene transcription or cell cycle progression. In this common motif, MAPK is activated by an earlier kinase in the cascade, called MAPK kinase, or MAPKK. Similarly, MAPKK is activated by MAPKK kinase, or MAPKKK. These kinases are sequentially phosphorylated when MAPKKK is activated, usually via a signal received by a membrane-bound receptor protein. MAPKKK activates MAPKK, and MAPKK activates MAPK.[41] Ultrasensitivity arises in this system due to several features:

  1. MAPK and MAPKK both require two separate phosphorylation events to be activated.
  2. The reversal of MAPK phosphorylation by specific phosphatases requires an increasing concentration of activation signals from each prior kinase to achieve an output of the same magnitude.
  3. The MAPKK is at a concentration above the KΜ for its specific phosphatase and MAPK is at a concentration above the KΜ for MAPKK.
Ultrasensitivity in the mitogen-activated protein (MAP) kinase cascade[41]

Besides the MAPK cascade, ultrasensitivity has also been reported in muscle glycolysis, in the phosphorylation of isocitrate dehydrogenase and in the activation of the calmodulin-dependent protein kinase II (CAMKII). [42]

An ultrasensitive switch has been engineered by combining a simple linear signaling protein (N-WASP) with one to five SH3 interaction modules that have autoinhibitory and cooperative properties. Addition of a single SH3 module created a switch that was activated in a linear fashion by exogenous SH3-binding peptide. Increasing number of domains increased ultrasensitivity. A construct with three SH3 modules was activated with an apparent Hill coefficient of 2.7 and a construct with five SH3 module was activated with an apparent Hill coefficient of 3.9. [43]

Translocation

During G2 phase of the cell cycle, Cdk1 and Cyclin B1 makes a complex and forms Maturation promoting factor(MPF). The complex accumulates in the nucleus due to phosphorylation of the Cyclin B1 at multiple sites, which inhibits nuclear export of the complex. Phosphorylation of Thr19 and Tyr15 residues of Cdc2 by Wee1 and MYT1 keeps the complex inactive and inhibits entry into mitosis whereas dephosphorylation of Cdc2 by CDC25C phosphatase at Thr19 and Tyr15 residues, activates the complex which is necessary in order to enter mitosis. Cdc25C phosphatase is present in the cytoplasm and in late G2 phase it is translocated in to nucleus by signaling such as PIK1[44], PIK3[45]. The regulated translocation and accumulation of the multiple required signaling cascade components; MPF and its activator Cdc25, in the nucleus generates efficient activation of the MPF and there by produces switch-like, ultrasensitive response to enter the mitosis.[4]

File:Ultrasensitive response mechanism.jpg
Increasing ultrasensitive response with increase in regulation of the localization of multiple components of the signaling cascade

The figure[4] shows different possible mechanisms for how increased regulation of the localization of signaling components by the stimulus (input signal) shifts the output from Michaelian response to ultrasensitive response. When stimulus is regulating only inhibition of Cdc2-CyclinB1 nuclear export, the outcome is Michaelian response, Fig (a). But if the stimulus can regulate localization of multiple components of the signaling cascade, i.e. inhibition of Cdc2-CyclinB1 nuclear export and translocation of the Cdc25C to nucleus, then the outcome is ultrasensitive response, Fig (b). As more components of the signaling cascade are regulated and localized by the stimulus, i.e. inhibition of Cdc2-CyclinB1 nuclear export, translocation of the Cdc25C to the nucleus and activation of Cdc25C; the output response becomes more and more ultrasensitive, Fig(c).[4]

Buffering (decoy)

During mitosis, mitotic spindle orientation is essential for determining the site of cleavage furrowing and position of daughter cells for subsequent cell fate determination[46]. This orientation is achieved by polarizing cortical factors and rapid alignment of the spindle with the polarity axis. Three cortical factors have been found to regulate the position of the spindle: heterotrimeric G protein α subunit (Gαi)[47], Partner of Inscuteable (Pins)[48], and Mushroom body defect (Mud)[49]. Gαi localizes at apical cortex to recruit Pins. Upon binding to GDP-bound Gαi, Pins is activated and recruits Mud to achieve polarized distribution of cortical factors[50]. N-terminal tetratricopeptide repeats (TPRs) in Pins is the binding region for Mud, but is autoinhibited by intrinsic C-terminal GoLoco domains (GLs) in the absence of of Gαi[51][52]. Activation of Pins by Gαi binding to GLs is highly ultrasensitive and is achieved through the following decoy mechanism[53]: GLs 1 and 2 act as a decoy domains, competing with the regulatory domain, GL3, for Gαi inputs. This intramolecular decoy mechanism allows Pins to establish it's of threshold and steepness in response to distinct Gαi concentration. At low Gαi inputs, the decoy GLs 1 and 2 are preferentially bound. At intermediate Gαi concentration, the decoys are nearly saturated, and GL3 begins to be populated. At higher Gαi concentration, the decoys are fully saturated and Gαi binds to GL3, leading to Pins activation. Ultrasensitivity of Pins in response to Gαi ensures that Pins is activated only at the apical cortex where Gαi concentration is above the threshold, allowing for maximal Mud recruitment.

Switching Behavior of GTPases

GTPases are enzymes capable of binding and hydrolyzing guanosine triphosphate (GTP). Small GTPases, such as Ran and Ras, can exist in either a GTP-bound form (active) or a GDP-bound form (inactive), and the conversion between these two forms grants them a switch-like behavior. [54] As such, small GTPases are involved in multiple cellular events, including nuclear translocation and signaling. [55] The transition between the active and inactive states is facilitated by guanine nucleotide exchange factors (GEFs) and GTPase activating proteins (GAPs).[56]

Computational studies on the switching behavior of GTPases have revealed that the GTPase-GAP-GEF system displays ultrasensitivity. [57] In their study, Lipshtat et al. simulated the effects of the levels of GEF and GAP activation on the Rap activation signaling network in response to signals from activated α2-adrenergic (α2R) receptors, which lead to degradation of the activated Rap GAP. They found that the switching behavior of Rap activation was ultrasensitive to changes in the concentration (i.e. amplitude) and the duration of the α2R signal, yielding Hill coefficients of nH=2.9 and nH=1.7, respectively (a Hill coefficient greater than nH=1 is characteristic of ultrasensitivity [58]). The authors confirmed this experimentally by treating neuroblasts with HU-210, which activates RAP through degradation of Rap GAP. Ultrasensitivity was observed both in a dose-dependent manner (nH=5±0.2), by treating cells with different HU-210 concentrations for a fixed time, and in a duration-dependent manner (nH=8.6±0.8), by treating cells with a fixed HU-210 concentration during varying times.

By further studying system, the authors determined that (the degree of responsiveness and ultrasensitivity) was heavily dependent on two parameters: the initial ratio of kGAP/kGEF, where the k’s incorporate both the concentration of active GAP or GEF and their corresponding kinetic rates; and the signal impact, which is the product of the degradation rate of activated GAP and either the signal amplitude or the signal duration. [59] The parameter kGAP/kGEF affects the steepness of the transition from the two states of the GTPase switch, with higher values (~10) leading to ultrasensitivity. The signal impact affects the switching point. Therefore, by depending on the ratio of concentrations rather than on individual concentrations, the switch-like behavior of the system can also be displayed outside of the zero-order regime.

Ultrasensitivity and Neuronal Potentiation

Persistent stimulation at the neuronal synapse can lead to markedly different outcomes for the post-synaptic neuron. Extended weak signaling can result in Long-Term Depression (LTD), in which activation of the post-synaptic neuron requires a stronger signal than before LTD was initiated. In contrast, Long-Term Potentiation (LTP) occurs when the post-synaptic neuron is subjected to a strong stimulus, and this results in strengthening of the neural synapse (i.e., less neurotransmitter signal is required for activation). In the CA1 region of the hippocampus, the decision between LTD and LTP is mediated solely by the level of intracellular Ca2+ at the post-synaptic dendritic spine. Low levels of Ca2+ (resulting from low-level stimulation) activates the protein phophatase calcineurin, which induces LTD. Higher levels of Ca2+ results in activation of the calmodulin-dependent protein kinase II (CaMKII), which leads to LTP. Interestingly, the difference in Ca2+ concentration required for a cell to undergo LTP is only marginally higher than that during LTD, and because neurons show bistability (either LTP or LTD) following persistent stimulation, this suggests that one or more components of the system respond in a switch-like, or ultrasensitive manner. Bradshaw et al., demonstrated that CaMKII (the LTP inducer) responds to intracellular calcium levels in an ultrasensitive manner, with <10% activity at 1.0 uM and ~90% activity at 1.5 uM, resulting in a Hill coefficient of ~8. Further experiments showed that this ultrasenstivity was mediated by co-operative binding of CaMKII by two molecules of Calmodulin (CaM), and autophosphorylation of activated CaMKII leading to a positive feedback loop[60]

In this way, intracellular calcium can induce a graded, non-ultrasensitive activation of calcineurin at a dynamic, low-level range intracellular calcium, leading to LTD, whereas the ultrasensitive activation of CaMKII results in a threshold of intracellular calcium levels, leading to a positive feedback loop that amplifies this signal and leads to a markedly different cellular outcome: LTP. Thus, binding of a single substrate to multiple enzymes with different sensitivities facilitates a bistable decision for the cell to undergo LTD or LTP.

Ultrasensitivity in Development

It has been suggested that zero-order ultrasensitivity may generate thresholds during development allowing for the conversion of a graded morphogen input to a binary switch-like response.[61] Melen et al. (2005) have found evidence for such a system in the patterning of the Drosophila embryonic ventral ectoderm.[62] In this system, graded mitogen activated protein kinase (MAPK) activity is converted to a binary output, the all-or-none degradation of the Yan transcriptional repressor. They found that MAPK phosphorylation of Yan is both essential and sufficient for Yan's degradation. Consistent with zero-order ultrasensitivity an increase in Yan protein lengthened the time required for degradation but had no effect on the border of Yan degradation in developing embryos. Their results are consistent with a situation where a large pool of Yan becomes either completely degraded or maintained. The particular response of each cell depends on whether or not the rate of reversible Yan phosphorylation by MAPK is greater or less than dephosphorylation. Thus, a small increase in MAPK phosphorylation can cause it to be the dominant process in the cell and lead to complete degradation of Yan.

Multistep-feedback loop mechanism also leads to ultrasensitivity

Multistep-feedback loop mechanism also leads to ultrasensitivity. There is paper introducing that engineering synthetic feedback loops using yeast mating mitogen-activated protein (MAP) kinase pathway as a model system.

In Yeast mating pathway: alpha-factor activates receptor, Ste2, and Ste4 and activated Ste4 recruits Ste5 complex to membrane, allowing PAK-like kinase Ste20 (membrane-localized) to activate MAPKKK Ste11. Ste11 and downstream kinases, Ste7 (MAPKK) and Fus3 (MAPK), are colocalized on the scaffold and activation of cascade leads to transcriptional program. They used pathway modulators outside of core cascade, Ste50 promotes activation of Ste11 by Ste20; Msg5 (negative, red) is MAPK phosphatase that deactivates Fus3 (Fig.2A).

What they built was circuit with enhanced ultrasensitive switch behavior by constitutively expressing a negative modulator, Msg5 which is one of MAPK phoaphatase and inducibly expressing a positive modulator, Ste50 which is pathway modulators outside of core cascade(Fig.2B). The success of this recruitment-based engineering strategy suggests that it may be possible to reprogram cellular responses with high precision. [63]

Flagellar motors and chemotaxis

The rotational direction of E. coli is controlled by the flagellar motor system. A ring of 34 FliM proteins around the rotor bind CheY, whose phosphorylation state determines whether the motor rotates in a clockwise or counterclockwise manner. The rapid switching mechanism is attributed to an ultrasensitive response, which has a Hill coefficient of ~10. This system has been proposed to follow a dissipative allosteric model, in which rotational switching is a result of both CheY binding and energy consumption from the proton motive force, which also powers the flagellar rotation[64].

References

  1. ^ . PMID 9572732. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  2. ^ . PMID 15907212. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  3. ^ . PMID 21562426. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  4. ^ a b c d e . PMID 9868363. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  5. ^ a b c Goldbeter, Albert; Koshland, Daniel E. "An Amplified Sensitivity Arising from Covalent Modification in Biological Systems". Proceedings of the National Academy of Sciences of the United States of America. JSTOR 11361. PMC 349147. PMID 6947258.
  6. ^ . PMID 21423715. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  7. ^ . PMID 18804166. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  8. ^ . PMID 16978046. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  9. ^ . PMID 18213378. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  10. ^ . doi:10.1088/1478-3975/8/4/046005. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  11. ^ . doi:10.1371/journal.pcbi.1002085. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)CS1 maint: unflagged free DOI (link)
  12. ^ . PMID 21451590. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  13. ^ . PMID 19438711. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  14. ^ . PMID 12023217. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  15. ^ . PMID 14744999. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  16. ^ : 540–9. doi:10.1016/j.molcel.2011.06.030. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  17. ^ : 1133–45. doi:10.1016/j.cell.2007.01.039. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  18. ^ {{cite journal}}: Empty citation (help)
  19. ^ a b . doi:10.1038/msb.2008.21. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  20. ^ . PMID 6501300. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  21. ^ . PMID 11408658. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  22. ^ . PMID 12023217. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  23. ^ . PMID 21354391. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  24. ^ . PMID 2418440. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  25. ^ a b c . PMID 18938177. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  26. ^ . PMID 19455136. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  27. ^ . PMID 18394191. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  28. ^ . PMID 9868363. {{cite journal}}: Cite journal requires |journal= (help); Missing or empty |title= (help)
  29. ^ a b Goldbeter, Albert (2005). "Zero-order switches and developmental thresholds" (PDF). Molecular Systems Biology. doi:doi:10.1038/msb4100042. {{cite journal}}: Check |doi= value (help)
  30. ^ Meinke, Marilyn H. (1986). "Zero-orde rultrasensitivity in the regulation of glycogen phosphorylase" (PDF). PNAS. 83: 2865–2868. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  31. ^ Goldbeter, Albert (1984). "Ultrasensitivity in Biochemical Systems Controlled by Covalent Modification". The Journal of Biological Chemistry. 259: 14441–14447. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  32. ^ a b Goulev, Youlian (2011). "Ultrasensitivity and Positive Feedback to Promote Sharp Mitotic Entry". Mol Cell. 41: 243–244. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help); line feed character in |title= at position 39 (help)
  33. ^ Trunnell, Nicole B. (2011). "Ultrasensitivity in the Regulation of Cdc25C by Cdk1". Mol Cell. 41: 263–274. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help); line feed character in |title= at position 35 (help)
  34. ^ Klein, Peter (30). "Mathematical Modeling Suggests Cooperative Interactions between a Disordered Polyvalent Ligand and a Single Receptor Site". Current Biology. 13: 1669–1678. doi:10.1016/j.cub.2003.09.027. PMID 14521832. Retrieved 11/18/2011. {{cite journal}}: Check date values in: |accessdate=, |date=, and |year= / |date= mismatch (help); Unknown parameter |coauthors= ignored (|author= suggested) (help); Unknown parameter |month= ignored (help)
  35. ^ Ravid, Tommer (2008). "Diversity of Degradation Signals in the Ubiquitin-Proteasome System". Nature Reviews Molecular Cell Biology. 9: 679–689. doi:doi:10.1038/nrm2468. PMID 18698327. Retrieved 11/18/2011. {{cite journal}}: Check |doi= value (help); Check date values in: |accessdate= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help); Unknown parameter |month= ignored (help)
  36. ^ Klein, Peter (30). "Mathematical Modeling Suggests Cooperative Interactions between a Disordered Polyvalent Ligand and a Single Receptor Site". Current Biology. 13: 1669–1678. doi:10.1016/j.cub.2003.09.027. PMID 14521832. Retrieved 11/18/2011. {{cite journal}}: Check date values in: |accessdate=, |date=, and |year= / |date= mismatch (help); Unknown parameter |coauthors= ignored (|author= suggested) (help); Unknown parameter |month= ignored (help)
  37. ^ Kõivomägi, Mardo (14). "Cascades of multisite phosphorylation control Sic1 destruction at the onset of S phase". Nature. doi:10.1038/nature10560. PMID 21993622. Retrieved November 18, 2011. {{cite journal}}: Check date values in: |date= and |year= / |date= mismatch (help); Unknown parameter |coauthors= ignored (|author= suggested) (help); Unknown parameter |month= ignored (help)
  38. ^ Dushek, O., van der Merwe, P. A., & Shahrezaei, V. (2011). Ultrasensitivity in multisite phosphorylation of membrane-anchored proteins Biophysical journal, 100(5), 1189–1197. doi:10.1016/j.bpj.2011.01.060 PMID 21354391
  39. ^ Dushek, O., van der Merwe, P. A., & Shahrezaei, V. (2011). Ultrasensitivity in multisite phosphorylation of membrane-anchored proteins Biophysical journal, 100(5), 1189–1197. doi:10.1016/j.bpj.2011.01.060 PMID 21354391
  40. ^ Bluethgen, Nils (2007). "Mechanisms Generating Ultrasensitivity, Bistability, and Oscillations in Signal Transduction(Introduction to Systems Biology)". Humana Press, Humboldt University Institute of Theoretical Biology Berlin Germany: 282–299. {{cite journal}}: Text "http://dx.doi.org/10.1007/978-1-59745-531-2_15" ignored (help)
  41. ^ a b c d Huang, Chi-Ying F.; James E. Ferrell, Jr. (September 1996). "Ultrasensitivity in the mitogen-activated protein kinase cascade". PNAS 93: 10078-10083. http://www.pnas.org/content/93/19/10078.full.pdf. Cite error: The named reference "Huang" was defined multiple times with different content (see the help page).
  42. ^ Bluethgen, Nils (2007). "Mechanisms Generating Ultrasensitivity, Bistability, and Oscillations in Signal Transduction(Introduction to Systems Biology)". Humana Press, Humboldt University Institute of Theoretical Biology Berlin Germany: 282–299. {{cite journal}}: Text "http://dx.doi.org/10.1007/978-1-59745-531-2_15" ignored (help)
  43. ^ Dueber, John (2007). "Engineering synthetic signaling proteins with ultrasensitive input/output control". Nature Biotechnology 25: 660–662. PMID 17515908. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  44. ^ Fumiko Toyoshima-Morimoto, Eri Taniguchi & Eisuke Nishida,"Plk1 promotes nuclear translocation of human Cdc25C during prophase ", EMBO reports 3, 4, 341–348 (2002)
  45. ^ Bahassi el M, Hennigan RF, Myer DL, Stambrook PJ.,"Cdc25C phosphorylation on serine 191 by Plk3 promotes its nuclear translocation.", Oncogene. 2004 Apr 8;23(15):2658-63.
  46. ^ Doe, C.Q. (2008). Neural stem cells: balancing self-renewal with differentiation. Development 135, 1575–1587.
  47. ^ Yu, F., Cai, Y., Kaushik, R., Yang, X., and Chia, W. (2003). Distinct roles of Galphai and Gbeta13F subunits of the heterotrimeric G protein complex in the mediation of Drosophila neuroblast asymmetric divisions. J. Cell Biol. 162, 623–633.
  48. ^ Izumi, Y., Ohta, N., Hisata, K., Raabe, T., and Matsuzaki, F. (2006). Drosophila Pins-binding protein Mud regulates spindle-polarity coupling and centrosome organization. Nat. Cell Biol. 8, 586–593.
  49. ^ Bowman, S.K., Neumuller, R.A., Novatchkova, M., Du, Q., and Knoblich, J.A. (2006). The Drosophila NuMA homolog Mud regulates spindle orientation in asymmetric cell division. Dev. Cell 10, 731–742.
  50. ^ Siller, K.H., Cabernard, C., and Doe, C.Q. (2006). The NuMA-related Mud protein binds Pins and regulates spindle orientation in Drosophila neuroblasts. Nat. Cell Biol. 8, 594–600.
  51. ^ Du, Q., and Macara, I.G. (2004). Mammalian Pins is a conformational switch that links NuMA to heterotrimeric G proteins. Cell 119, 503–516.
  52. ^ Nipper, R.W., Siller, K.H., Smith, N.R., Doe, C.Q., and Prehoda, K.E. (2007). Galphai generates multiple Pins activation states to link cortical polarity and spindle orientation in Drosophila neuroblasts. Proc. Natl. Acad. Sci. USA 104, 14306–14311.
  53. ^ Smith, N. R., and Prehoda, K. E. (2011). Robust spindle alignment in Drosophila neuroblasts by ultrasensitive activation of pins. Mol Cell 43, 540–549.
  54. ^ Yang, Z (2002). "Small GTPases". Plant Cell. 14: s375–s388.
  55. ^ Heider, D (2010). "Insights into the classification of small GTPases". Advances and Applications in Bioinformatics and Chemistry. 3: 15–24. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  56. ^ Bourne, HR (1991). "The GTPase superfamily: conserved structure and molecular mechanism". Nature. 349: 117–127. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  57. ^ Lipshtat, A (2010). "Design of versatile biochemical switches that respond to amplitude, duration, and spatial cues". PNAS. 107(3): 1247–1252. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  58. ^ Ferrel, JE Jr. (1999). "Building a cellular switch: more lessons from a good egg". Bioessays. 21: 866–870.
  59. ^ Lipshtat, A (2010). "Design of versatile biochemical switches that respond to amplitude, duration, and spatial cues". PNAS. 107(3): 1247–1252. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  60. ^ Bradshaw JM, Kubota Y, Meyer T, Schulman H. (2003) "An ultrasensitive Ca2+/calmodulin-dependent protein kinase II-protein phosphatase 1 switch facilitates specificity in postsynaptic calcium signaling" PNAS 100(18):10512-7
  61. ^ Goldbeter A, Wolpert L (1990) Covalent modification of proteins as a threshold mechanism in development. J Theor Biol 142: 243–250
  62. ^ Melen et al., 2005 G.J. Melen, S. Levy, N. Barkai and B.Z. Shilo, Threshold responses to morphogen gradients by zero-order ultrasensitivity. Mol. Syst. Biol., 1 (2005), p. 0028.
  63. ^ Caleb J. Bashor, Noah C. Helman, Shude Yan, Wendell A. Lim, “Using Engineered Scaffold Interactions to Reshape MAP Kinase Pathway Signaling Dynamics” Science 2008
  64. ^ Tu, Y. (2008) "The nonequilibrium mechanism for ultrasensitivity in a biological switch: Sensing by Maxwell's demons." "PNAS" 105 (33) 11737-11741.