Difference between revisions of "Team:ETH Zurich/Circuit/Fd Heat Sensor"

 
(22 intermediate revisions by 3 users not shown)
Line 1: Line 1:
 
{{ETH_Zurich/Head_N_Prefix}}
 
{{ETH_Zurich/Head_N_Prefix}}
 
<html>
 
<html>
<link rel="stylesheet" type="text/css" href="/Template:ETH_Zurich/background_css?action=raw&ctype=text/css">
+
<!--
 +
<link rel="stylesheet" type="text/css" href="/Template:ETH_Zurich/heat_sensor_css?action=raw&ctype=text/css">
 +
-->
 
</html>  
 
</html>  
 
{{ETH_Zurich/Head_N_Suffix}}
 
{{ETH_Zurich/Head_N_Suffix}}
Line 7: Line 9:
 
<html>
 
<html>
 
<main role="main">
 
<main role="main">
<h1 class="headline">Heat Sensor</h1>
+
<h1 class="headline">Function D: Heat Sensor</h1>
  
<section>
+
<p><em>This is a detailed description of an individual function of our circuit. To access other functions and get an overview of the whole circuit, visit the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit">Circuit</a> page.</em></p>
<h1>Introduction</h1>
+
  
<p>Sensing the temperature is crucial for the lifecycle of many bacteria and viruses. We adapt this function from a specific heat-inducible system and engineer it to fit our needs for activating the toxin release in CATE. </p>
+
<section class="first">
 +
    <h1>Introduction</h1>
 +
    <p>Sensing the temperature is crucial for the lifecycle of many bacteria and viruses. <a href="#bib1" class="forward-ref">[1]</a>
  
<p>
 
Various thermosensitive operator systems exist in nature and they differ massively in their way of function. Four general classes of thermosensors exist: DNA, RNA, protein or lipid-protein thermosensors.
 
  
 +
We adapt this function from a specific heat-inducible system and engineer it to fit our needs for activating the toxin release in CATE.</p>
 +
</section>
  
<p>
+
<section>
 +
    <h1>Different heat sensors</h1>
 +
    <p>Various thermosensitive operator systems exist in nature and they differ massively in their way of function.
  
<img src="https://static.igem.org/mediawiki/2017/2/21/T--ETH_Zurich--DNA_Heat_Sensors.png" alt="DNA based heat sensor">
+
Four general classes of thermosensors exist: DNA, RNA, protein or lipid-protein thermosensors. <a href="#bib2" class="forward-ref">[2]</a></p>
</p>
+
    <p>DNA thermosensors rely on the bending of DNA at lower temperatures, which enables cooperative binding of DNA-associated proteins (Figure 1). At lower temperatures, the DNA binding proteins can cover the operator region and prevent transcription. At higher temperatures, the DNA unbends and the operator region can promote transcription of the downstream genes.</p>
 +
    <figure class="fig-nonfloat" style="width:500px;">
 +
        <img alt="RNA based heat sensor"
 +
        src="https://static.igem.org/mediawiki/2017/2/21/T--ETH_Zurich--DNA_Heat_Sensors.png"/>
 +
    <figcaption>Figure 1. DNA-based thermosensor.</figcaption>
 +
    </figure>
  
DNA thermosensors rely on the bending of DNA at lower temperatures, which enables cooperative binding of DNA-associated proteins. At lower temperatures, the DNA binding proteins can cover the operator region and prevent transcription. At higher temperatures, the DNA unbends and the operator region can promote transcription of the downstream genes.</p>
+
    <p>RNA based thermosensors form a stem-loop in the messenger RNA, which hides the Shine-Dalgarno sequence and the AUG translation initiation codon (Figure 2). At higher temperatures, the hydrogen bonds of the stem loop break apart and the ribosomal subunits can associate with the RNA and initiate translation.</p>
  
<p>
+
    <figure class="fig-nonfloat" style="width:500px;">
 +
        <img src="https://static.igem.org/mediawiki/2017/9/90/--ETH_Zurich--RNA_Heat_Sensors.png" alt="RNA based heat sensor">
 +
    <figcaption>Figure 2. RNA-based thermosensor.</figcaption>
 +
    </figure>
 +
    <p>RNA thermosensors can act without a translated mediator protein. In contrast, for protein-based temperature sensing, a protein needs to be expressed and available in the cytoplasm (Figure 3). This comes with the potential to modulate the sensors characteristics by modulating the protein expression levels. This can be achieved by modifying the RBS sequence to generate small and easily screenable RBS libraries. <a href="#bib3" class="forward-ref">[3]</a>
  
<img src="https://static.igem.org/mediawiki/2017/9/90/--ETH_Zurich--RNA_Heat_Sensors.png" alt="RNA based heat sensor">
+
This is a reason why we use a protein based thermosensor.</p>
</p>
+
  
<p>
+
    <figure class="fig-nonfloat" style="width:500px;">
RNA based thermosensors form a stem-loop in the messenger RNA, which hides the Shine-Dalgarno sequence and the AUG translation initiation codon. At higher temperatures, the hydrogen bonds of the stem loop break apart and the ribosomal subunits can associate with the RNA and initiate translation. While DNA- and RNA thermosensors act before translation, for protein mediated temperature sensing, a translated protein needs to be present in the cytoplasm.</p>
+
        <img alt="Protein based heat sensor"
 +
        src="https://static.igem.org/mediawiki/2017/b/bb/--ETH_Zurich--Protein_Heat_Sensors.png"/>
 +
    <figcaption>Figure 3. Protein-based thermosensor.</figcaption>
 +
    </figure>
  
<p>
+
   
 +
</section>
  
<p>
+
<section>
 +
    <h1>CATE's heat sensor</h1>
 +
    <p>For CATE, we chose the TlpA thermosensor, which is derived from <span class="bacterium">Salmonella</span> and belongs to the protein thermosensors. <a href="#bib4" class="forward-ref">[4]</a> The advantage of this system is the high on/off-ratio of up to 355 <a href="#bib4" class="forward-ref">[4]</a> and induction temperature of 45 °C, a temperature not reached by fever, but still below levels that cause damage in tissues surrounding a tumor. We adapt this function so that we can use it to trigger the release of the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fc_Anti_Cancer_Toxin">Anti-Cancer Toxin</a>.</p>
  
<img src="https://static.igem.org/mediawiki/2017/b/bb/--ETH_Zurich--Protein_Heat_Sensors.png" alt="Protein based heat sensor">
+
    <p>The TlpA system consists of a constitutively expressed regulator protein called TlpA and a promotor containing a binding site for TlpA called P<sub>TlpA</sub> (Figure 4). TlpA contains an approximately 300-residue coiled-coil domain at the C-terminus that uncoils between 42 °C and 45 °C. In low temperatures, its N-terminal domain is in a dimeric state and can bind the 52-bp P<sub>TlpA</sub>. Transcription of the downstream gene can therefore happen at temperatures above 42 °C but not below. <a href="#bib4" class="forward-ref">[4]</a> <a href="#bib5" class="forward-ref">[5]</a></a></p>
</p>
+
  
This comes with the potential to tune the translation initiation rate by changing the ribosome binding site affinity to the ribosome and is a reason why we use a protein based thermosensor.</p>
+
    <figure class="fig-nonfloat" style="width:800px;">
 +
        <img alt="TlpA heat sensor"
 +
        src="https://static.igem.org/mediawiki/2017/7/7e/T--ETH_Zurich--Heat_Sensor_Function.png"/>
 +
    <figcaption>Figure 4. CATE's protein-based thermosensor. Protein E is regulated by TlpA-mediated thermosensing in the circuit of CATE.</figcaption>
 +
    </figure>
 
</section>
 
</section>
 
  
 
<section>
 
<section>
<h1>CATE's heat sensor</h1>
+
    <h1>The heat sensor is the safety Checkpoint 2</h1>
<p>
+
    <p>In a cancer treatment with CATE, the bacteria will prepare the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fc_Anti_Cancer_Toxin">Anti-Cancer Toxin</a> in the cytoplasm together with the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fb_MRI_Contrast_Agent">MRI Contrast Agent</a> as soon as they sense a tumor environment with the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fa_Tumor_Sensor">Tumor Sensor</a>. At this point the patient is analyzed in an MRI and a doctor can decide if the cancer toxin should be released. This is an additional safety checkpoint that is unprecedented for systemically (to the whole body) applied cancer therapeutics.</p>
For CATE, we chose the TlpA thermosensor, which is derived from <span class="bacterium">Salmonella</span> and belongs to the protein thermosensors. The advantage of this system is the high on/off-ratio of up to 300 and induction temperature of 45 °C, a temperature not reached by fever, but still below levels that cause damage in tissues surrounding a tumor. We adapt this function so that we can use it to trigger the release of the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fc_Anti_Cancer_Toxin">Anti-Cancer Toxin</a>.
+
</p>
+
  
<p>
+
    <p>If the physician decides to proceed with the treatment, he can use focused ultrasound, a technology already included in modern MRI devices. It can heat up patients tissue in a precisely defined area inside the body. Focused ultrasound is normally used with very high teperatures to destroy tumors. However, the strong heat diffuses out of the tumor and thus damages surrounding healthy tissue. To prevent this off-target damage, we use the same focused ultrasound technology at a lower temperature, just high enough for CATE to be able to sense a change. The doctor will focus the ultrasound waves with the device to the area where the tumor is and heat the tumor tissue to 45 °C for up to 3 h (Figure 5). The heat sensor notices this change in temperature by unfolding the coiled-coil domains of the TlpA dimers. As a consequence, the dimers fall apart to monomers, which will not bind the P<sub>TlpA</sub> anymore. The P<sub>TlpA</sub> in an unbound state has a strong promotor activity and induces expression of protein E, which leads to <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fe_Cell_Lysis">Cell Lysis</a>.
The TlpA system consists of a constitutively expressed regulator protein called TlpA and an inducible TlpA operator-promotor called P<sub>TlpA</sub>. TlpA contains an approximately 300-residue coiled-coil domain at the C-terminus that uncoils between 42 °C and 45 °C. In low temperatures, its N-terminal domain is in a dimeric state and can bind the 52-bp P<sub>TlpA</sub>. Transcription of the downstream gene can therefore happen at temperatures above 42 °C but not below. <a href="#bib1" class="forward-ref">[1]</a><a href="#bib2" class="forward-ref">[2]</a>
+
 
</p>
+
    <figure class="fig-nonfloat" style="width:800px;">
 +
        <img alt="Release of azurin"
 +
        src="https://static.igem.org/mediawiki/2017/5/5d/T--ETH_Zurich--Circuit_Azu_Lysis.png">
 +
        <figcaption>Figure 5. The safety checkpoint 2. TlpA dimers bind P<sub>TlpA</sub> and the accumulated Anti-Cancer Toxin (azurin) stays inside the cell. A pulse of focused ultrasound increases the temperature to 45 °C and unfolds the TlpA dimers. Unfolded TlpA does not bind the P<sub>TlpA</sub> anymore. Unbound P<sub>TlpA</sub> is a strong promoter and leads to protein E expression. Protein E interferes with cell wall synthesis and leads to Cell Lysis. Previously accumulated azurin is released to the extracellular space during lysis of CATE. </figcaption>
 +
    </figure>
  
 
<p>
 
<p>
<img src="https://static.igem.org/mediawiki/2017/7/7e/T--ETH_Zurich--Heat_Sensor_Function.png" alt="TlpA heat sensor">
+
Following cell lysis, azurin is released to the extracellular space (Figure 5).
 
+
 
</p>
 
</p>
 
</section>
 
</section>
 
<section>
 
<h1>Checkpoint 2</h1>
 
<p>The Heat Sensor is used as Checkpoint 2 to ensure that only bacteria in the tumor site release the <a href="https://2017.igem.org/Team:ETH_Zurich/Circuit/Fc_Anti_Cancer_Toxin">Anti-Cancer Toxin</a>.</p>
 
</section>
 
 
  
 
<section class="references">
 
<section class="references">
 
     <h1>References</h1>
 
     <h1>References</h1>
 
     <ol>
 
     <ol>
         <li id="bib1"><a href="#ref1">^ </a>Hurme, R., Berndt, K.D., Namork, E. & Rhen, M. "DNA binding exerted by a bacterial gene regulator with an extensive coiled-coil domain." <i>J. Biol. Chem.</i> 271 (1996): 12626–12631.</li>
+
         <li id="bib1">Sengupta, Piali, and Paul Garrity. "Sensing temperature." <cite>Current Biology</cite> 23.8 (2013): R304-R307.<a href="https://doi.org/10.1016/j.cub.2013.03.009"> doi: 10.1016/j.cub.2013.03.009</a></li>
 
+
         <li id="bib2">Hurme, Reini, and Mikael Rhen. "Temperature sensing in bacterial gene regulation—what it all boils down to." <cite>Molecular microbiology</cite> 30.1 (1998): 1-6.<a href="https://doi.org/10.1046/j.1365-2958.1998.01049.x"> doi: 10.1046/j.1365-2958.1998.01049.x</a></li>
         <li id="bib2"><a href="#ref2">^ </a>Piraner, Dan I., et al. "Tunable thermal bioswitches for in vivo control of microbial therapeutics."<i>Nature chemical biology</i> 13.1 (2017): 75-80.</li>
+
  <li id="bib3">Jeschek, Markus, Daniel Gerngross, and Sven Panke. "Rationally reduced libraries for combinatorial pathway optimization minimizing experimental effort." <cite>Nature communications</cite> 7 (2016).<a href="https://doi.org/10.1038/ncomms11163"> doi: 10.1038/ncomms11163</a></li>
 
+
<li id="bib4">Hurme, Reini, et al. "DNA binding exerted by a bacterial gene regulator with an extensive coiled-coil domain." <cite>Journal of Biological Chemistry</cite> 271.21 (1996): 12626-12631.<a href="https://doi.org/10.1074/jbc.271.21.12626"> doi: 10.1074/jbc.271.21.12626</a></li>
 
+
<li id="bib5">Piraner, Dan I., et al. "Tunable thermal bioswitches for in vivo control of microbial therapeutics." <cite>Nature chemical biology</cite> 13.1 (2017): 75-80.<a href="https://doi.org/10.1038/nchembio.2233"> doi: 10.1038/nchembio.2233</a></li>
<!--
+
        <li id="bib3"><a href="#ref3">^ </a>Din, M. Omar, et al. "Synchronized cycles of bacterial lysis for in vivo delivery." <i>Nature</i> 536.7614 (2016): 81-85.</li>
+
        <li id="bib4"><a href="#ref4">^ </a>Roof, William D., and R. Young. "Phi X174 E complements lambda S and R dysfunction for host cell lysis." <i>Journal of bacteriology</i> 175.12 (1993): 3909-3912.</li>
+
<li id="bib5"><a href="#ref5">^ </a>Lubitz, W., R. E. Harkness, and E. E. Ishiguro. "Requirement for a functional host cell autolytic enzyme system for lysis of Escherichia coli by bacteriophage phi X174." <i>Journal of bacteriology</i> 159.1 (1984): 385-387.</i></li>
+
<li id="bib6"><a href="#ref6">^ </a>Witte, Angela, et al. "Endogenous transmembrane tunnel formation mediated by phi X174 lysis protein E." <i>Journal of bacteriology</i> 172.7 (1990): 4109-4114.</li>
+
    </ol>
+
-->
+
  
 +
</ol>
 
</section>
 
</section>
 
</main>
 
</main>
 
</html>
 
</html>
 
{{ETH_Zurich/Footer_N}}
 
{{ETH_Zurich/Footer_N}}

Latest revision as of 00:06, 2 November 2017

Function D: Heat Sensor

This is a detailed description of an individual function of our circuit. To access other functions and get an overview of the whole circuit, visit the Circuit page.

Introduction

Sensing the temperature is crucial for the lifecycle of many bacteria and viruses. [1] We adapt this function from a specific heat-inducible system and engineer it to fit our needs for activating the toxin release in CATE.

Different heat sensors

Various thermosensitive operator systems exist in nature and they differ massively in their way of function. Four general classes of thermosensors exist: DNA, RNA, protein or lipid-protein thermosensors. [2]

DNA thermosensors rely on the bending of DNA at lower temperatures, which enables cooperative binding of DNA-associated proteins (Figure 1). At lower temperatures, the DNA binding proteins can cover the operator region and prevent transcription. At higher temperatures, the DNA unbends and the operator region can promote transcription of the downstream genes.

RNA based heat sensor
Figure 1. DNA-based thermosensor.

RNA based thermosensors form a stem-loop in the messenger RNA, which hides the Shine-Dalgarno sequence and the AUG translation initiation codon (Figure 2). At higher temperatures, the hydrogen bonds of the stem loop break apart and the ribosomal subunits can associate with the RNA and initiate translation.

RNA based heat sensor
Figure 2. RNA-based thermosensor.

RNA thermosensors can act without a translated mediator protein. In contrast, for protein-based temperature sensing, a protein needs to be expressed and available in the cytoplasm (Figure 3). This comes with the potential to modulate the sensors characteristics by modulating the protein expression levels. This can be achieved by modifying the RBS sequence to generate small and easily screenable RBS libraries. [3] This is a reason why we use a protein based thermosensor.

Protein based heat sensor
Figure 3. Protein-based thermosensor.

CATE's heat sensor

For CATE, we chose the TlpA thermosensor, which is derived from Salmonella and belongs to the protein thermosensors. [4] The advantage of this system is the high on/off-ratio of up to 355 [4] and induction temperature of 45 °C, a temperature not reached by fever, but still below levels that cause damage in tissues surrounding a tumor. We adapt this function so that we can use it to trigger the release of the Anti-Cancer Toxin.

The TlpA system consists of a constitutively expressed regulator protein called TlpA and a promotor containing a binding site for TlpA called PTlpA (Figure 4). TlpA contains an approximately 300-residue coiled-coil domain at the C-terminus that uncoils between 42 °C and 45 °C. In low temperatures, its N-terminal domain is in a dimeric state and can bind the 52-bp PTlpA. Transcription of the downstream gene can therefore happen at temperatures above 42 °C but not below. [4] [5]

TlpA heat sensor
Figure 4. CATE's protein-based thermosensor. Protein E is regulated by TlpA-mediated thermosensing in the circuit of CATE.

The heat sensor is the safety Checkpoint 2

In a cancer treatment with CATE, the bacteria will prepare the Anti-Cancer Toxin in the cytoplasm together with the MRI Contrast Agent as soon as they sense a tumor environment with the Tumor Sensor. At this point the patient is analyzed in an MRI and a doctor can decide if the cancer toxin should be released. This is an additional safety checkpoint that is unprecedented for systemically (to the whole body) applied cancer therapeutics.

If the physician decides to proceed with the treatment, he can use focused ultrasound, a technology already included in modern MRI devices. It can heat up patients tissue in a precisely defined area inside the body. Focused ultrasound is normally used with very high teperatures to destroy tumors. However, the strong heat diffuses out of the tumor and thus damages surrounding healthy tissue. To prevent this off-target damage, we use the same focused ultrasound technology at a lower temperature, just high enough for CATE to be able to sense a change. The doctor will focus the ultrasound waves with the device to the area where the tumor is and heat the tumor tissue to 45 °C for up to 3 h (Figure 5). The heat sensor notices this change in temperature by unfolding the coiled-coil domains of the TlpA dimers. As a consequence, the dimers fall apart to monomers, which will not bind the PTlpA anymore. The PTlpA in an unbound state has a strong promotor activity and induces expression of protein E, which leads to Cell Lysis.

Release of azurin
Figure 5. The safety checkpoint 2. TlpA dimers bind PTlpA and the accumulated Anti-Cancer Toxin (azurin) stays inside the cell. A pulse of focused ultrasound increases the temperature to 45 °C and unfolds the TlpA dimers. Unfolded TlpA does not bind the PTlpA anymore. Unbound PTlpA is a strong promoter and leads to protein E expression. Protein E interferes with cell wall synthesis and leads to Cell Lysis. Previously accumulated azurin is released to the extracellular space during lysis of CATE.

Following cell lysis, azurin is released to the extracellular space (Figure 5).

References

  1. Sengupta, Piali, and Paul Garrity. "Sensing temperature." Current Biology 23.8 (2013): R304-R307. doi: 10.1016/j.cub.2013.03.009
  2. Hurme, Reini, and Mikael Rhen. "Temperature sensing in bacterial gene regulation—what it all boils down to." Molecular microbiology 30.1 (1998): 1-6. doi: 10.1046/j.1365-2958.1998.01049.x
  3. Jeschek, Markus, Daniel Gerngross, and Sven Panke. "Rationally reduced libraries for combinatorial pathway optimization minimizing experimental effort." Nature communications 7 (2016). doi: 10.1038/ncomms11163
  4. Hurme, Reini, et al. "DNA binding exerted by a bacterial gene regulator with an extensive coiled-coil domain." Journal of Biological Chemistry 271.21 (1996): 12626-12631. doi: 10.1074/jbc.271.21.12626
  5. Piraner, Dan I., et al. "Tunable thermal bioswitches for in vivo control of microbial therapeutics." Nature chemical biology 13.1 (2017): 75-80. doi: 10.1038/nchembio.2233