Δευτέρα 20 Απριλίου 2020

Role of Molecular Chaperones in Carcinogenesis: Mechanism, Diagnosis, and Treatment

Role of Molecular Chaperones in Carcinogenesis: Mechanism, Diagnosis, and Treatment:

jo.banner.jpgThe role of molecular chaperones in carcinogenesis is a central theme of many current research efforts worldwide. It is pertinent to clarify that while various chaperones are Hsps, many are not, and vice versa not all Hsps are chaperones. An example of the latter is Hsp32, better known as heme oxygenase-1 (HO-1), which is dealt with in one of the articles in this Special Issue because of its probable involvement in certain types of carcinogenesis. Likewise, long noncoding RNAs are currently emerging as modulators of metastasization, and one contribution deals with this interesting new issue in Oncology.


While the involvement of chaperones in cancer progression has been extensively reported, the mechanism of their participation in carcinogenesis is largely unknown. Nevertheless, extant information is enough for considering these molecules as a promising biomarker for diagnosis and patient monitoring in some types of cancer and also promising to be used as therapeutic targets or agents. The aim of this Special Issue is to gather information about these themes.

Characterization of mechanisms underlying the role of molecular chaperones in cancer is an emerging issue in the oncology field. Molecular chaperones are involved in many biochemical pathways essential for tumor cell survival, and this makes them candidates for consideration as key players in the biochemistry of cancer. Therefore, the development of tools for diagnosis and treatment targeting chaperones is currently an active discipline within oncology.

Some malignant tumors can be classified as “chaperonopathies by mistake” because the molecular chaperones in the cancer cells contribute to their proliferation and mediate their resistance against antitumor defenses and facilitate metastasization. Chaperones are thus helping the “enemy” so to speak and are therefore “mistaken.” These chaperones work for the tumor rather than to defend the host. Efforts must be directed toward finding ways to eliminate or block these “mistaken” chaperones. This strategy of negative chaperonotherapy is currently being incorporated to the battery of other approaches such as chemo- and immunotherapy to treat cancer.

In this Special Issue, various examples of tumors in which Hsp-chaperones and a Hsp that is not a chaperone play a noteworthy role are discussed along with the potential of chaperonotherapy.

The review by Das et al. highlights recent advances and perspectives in Hsp-based cancer immunotherapy. The importance of research on the role of Hsp-chaperones, specifically Hsp27, Hsp60, Hsp70, and Hsp90, in modulating carcinogenesis is discussed, emphasizing their critical role in determining the balance between protective and destructive immunological responses within the tumor microenvironment; the possibilities involving Hsp27, Hsp70, and Hsp90 are clearly schematized in the first figure of the article, while pathways of presentation of tumor antigens by Hsps to antigen-presenting cells (APCs) are summarized in the second figure. Along this line of thought, Hsps are seen as an effective therapeutic option for some malignant tumors, including melanoma. Recent progress in this field is discussed, including anticancer vaccines, specifically those based on Hsp70 and Hsp90 are summarized in Table 1 of the article. These vaccines have been shown to be active against a spectrum of tumor antigens since they induce T-lymphocyte activation as well as stimulation of antigens uptake by APCs.

In the review by Fucarino et al., emphasis is given to the study of the molecular mechanisms in carcinogenesis that involve chaperones. Specifically, they focus on the chaperonin Hsp60 alone or in complex with Hsp10, and its implication in lung cancer, and also analyze the broad set of Hsp60 interactors, some of which are listed in Table 1 of the article. Noteworthy is the relationship of the fragile histidine triad (FHIT) protein, a tumor-suppressor factor, with Hsp60 and Hsp10 in complex, in mitochondria. FHIT, like many other mitochondrial proteins, depends for its correct folding on the Hsp60/Hsp10 chaperoning complex; it is, therefore, possible that a chaperonopathy affecting either one or the other of these two chaperonins will result in a decrease of tumor suppression, namely, will favor tumor growth. This is certainly a topic for future investigation because it may help find ways to apply positive chaperonotherapy to oppose tumor progression by administering Hsp60 and/or Hsp10 or boosting their activities, thereby reenforcing the chaperoning of FHIT protein correct folding, which will thus become an effective tumor suppressor. Continuous stress conditions in the respiratory mucosa, such as cigarette smoking, increase Hsp60 chaperonin levels outside mitochondria. High levels of this chaperonin are associated with tumor deterioration in some cancer types, but in other types, the reverse is seen, namely, tumor progression is favored probably due to inhibition of apoptosis and senescence in tumor cells by the chaperonin.

The paper by Wu et al. examines the effects of long noncoding RNA lnc-TLN2-4:1 in gastric cancer (GC). Forty-nine patients were recruited for this study who had not been submitted to chemotherapy before and were followed for four years. The authors demonstrate that lnc-TLN2-4:1 is decreased in GC tissue compared with matched normal tissue and is involved in poor overall survival rates of GC patients. Moreover, they show that lnc-TLN2-4:1 overexpression inhibits GC cell migration and invasion, but does not affect GC cell proliferation, suggesting its involvement as a tumor suppressor of GC metastases. In conclusion, Inc-TLN2-4:1 is a suppressor of metastasization and when it is under-regulated like in GC the survival rate of patients is poor.

The paper by Castruccio Castracani et al. presents data on the effect of estradiol E2 in human glioblastoma-multiforme (GBM) cells, focusing on proliferative capacity and mitochondrial functions. Expression of genes involved in mitochondrial biogenesis, oxidative phosphorylation, and dynamics was measured. Also, nuclear translocation of Nrf2 was assessed. The results showed that E2 increases the proliferation of glioblastoma cells and changes the expression of various genes among those investigated, e.g., those involved in oxidative phosphorylation. E2 also increased nuclear translocation of Nrf2 resulting in the induction of one of its target genes, hemeoxygenase-1, which is associated with the increase of both, chemoresistance and tumor-cell proliferation. These data show that E2 induces GBM proliferation and enhances its mitochondrial physiology, both of which certainly play a role in the well-known resistance of this tumor to all kinds of therapies, and open new avenues for developing novel treatment strategies.

In the paper by Kam et al., the authors report on the compound 2-methoxyestradiol as a potential anticancer agent, using as a model the A375 melanoma cells. They also studied the effect of polyphenol ferulic acid using the same cells. By applying MTT, flow cytometry, and Western blotting, the authors examined the molecular mechanisms underpinning the compounds’ actions. These are partly related to the reduction of Hsp60 and Hsp90 levels and the induction of nitric oxide in A375 melanoma cells. The authors did not observe any changes in Hsp70 levels after 2-methoxyestradiol and ferulic acid treatment separately or in combination. This information is pertinent to evaluate chemoresistance mechanisms since Hsp70 accumulation reduces induction of cancer-cell death and decreases the antitumor efficacy of some therapeutic agents.

Conflicts of Interest
The editors declare that they have no conflicts of interest regarding the publication of this Special Issue.

Acknowledgments
The authors would like to thank the authors for their contributions to this Special Issue. ECdeM was partially supported by IMET (IMET contribution number: IMET 20-006).

Everly Conway de Macario
Alessandro Pitruzzella
Agata Grazia D’Amico

Copyright
Copyright © 2020 Everly Conway de Macario et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.


Chaperone (protein)

From Wikipedia, the free encyclopedia
Jump to navigationJump to search
A top-view of the GroES/GroEL bacterial chaperone complex model
In molecular biologymolecular chaperones are proteins that assist the conformational folding or unfolding and the assembly or disassembly of other macromolecular structures. Chaperones are present when the macromolecules perform their normal biological functions and have correctly completed the processes of folding and/or assembly. The chaperones are concerned primarily with protein folding. The first protein to be called a chaperone assists the assembly of nucleosomes from folded histones and DNA and such assembly chaperones, especially in the nucleus,[1][2] are concerned with the assembly of folded subunits into oligomeric structures.[3]
One major function of chaperones is to prevent both newly synthesised polypeptide chains and assembled subunits from aggregating into nonfunctional structures. It is for this reason that many chaperones, but by no means all, are heat shock proteins because the tendency to aggregate increases as proteins are denatured by stress. In this case, chaperones do not convey any additional steric information required for proteins to fold. However, some highly specific 'steric chaperones' do convey unique structural (steric) information onto proteins, which cannot be folded spontaneously. Such proteins violate Anfinsen's dogma,[4] requiring protein dynamics to fold correctly.
Various approaches have been applied to study the structure, dynamics and functioning of chaperones. Bulk biochemical measurements have informed us on the protein folding efficiency, and prevention of aggregation when chaperones are present during protein folding. Recent advances in single-molecule analysis[5] have brought insights into structural heterogeneity of chaperones, folding intermediates and affinity of chaperones for unstructured and structured protein chains.

Location and functions[edit]

Some chaperone systems work as foldases: they support the folding of proteins in an ATP-dependent manner (for example, the GroEL/GroES or the DnaK/DnaJ/GrpE system). Although most newly synthesized proteins can fold in absence of chaperones, a minority strictly requires them for the same. Other chaperones work as holdases: they bind folding intermediates to prevent their aggregation, for example DnaJ or Hsp33.[6] Chaperones can also work as disaggregases, i.e. they can interact with aberrant protein assemblies and revent them to monomers.[7] Some chaperones can assist in protein degradation, leading proteins to protease systems, such as the ubiquitin-proteasome system in eukaryotes.[8]
Many chaperones are heat shock proteins, that is, proteins expressed in response to elevated temperatures or other cellular stresses.[9] The reason for this behaviour is that protein folding is severely affected by heat and, therefore, some chaperones act to prevent or correct damage caused by misfolding.
Macromolecular crowding may be important in chaperone function. The crowded environment of the cytosol can accelerate the folding process, since a compact folded protein will occupy less volume than an unfolded protein chain.[10] However, crowding can reduce the yield of correctly folded protein by increasing protein aggregation.[11][12] Crowding may also increase the effectiveness of the chaperone proteins such as GroEL,[13] which could counteract this reduction in folding efficiency.[14]
More information on the various types and mechanisms of a subset of chaperones that encapsulate their folding substrates (e.g. GroES) can be found in the article for chaperonins. Chaperonins are characterized by a stacked double-ring structure and are found in prokaryotes, in the cytosol of eukaryotes, and in mitochondria.
Other types of chaperones are involved in transport across membranes, for example membranes of the mitochondria and endoplasmic reticulum (ER) in eukaryotes. A bacterial translocation—specific chaperone maintains newly synthesized precursor polypeptide chains in a translocation-competent (generally unfolded) state and guides them to the translocon.[15]
New functions for chaperones continue to be discovered, such as bacterial adhesin activity, induction of aggregation towards non-amyloid aggregates,[16] suppression of toxic protein oligomers via their clustering,[17][18] and in responding to diseases linked to protein aggregation[19] (e.g. see prion) and cancer maintenance.[20]

Human chaperone proteins[edit]

Chaperones are found in, for example, the endoplasmic reticulum (ER), since protein synthesis often occurs in this area.

Endoplasmic reticulum[edit]

In the endoplasmic reticulum (ER) there are general, lectin- and non-classical molecular chaperones helping to fold proteins.

Nomenclature and examples of bacterial and archaeal chaperones[edit]

There are many different families of chaperones; each family acts to aid protein folding in a different way. In bacteria like E. coli, many of these proteins are highly expressed under conditions of high stress, for example, when the bacterium is placed in high temperatures. For this reason, the term "heat shock protein" has historically been used to name these chaperones. The prefix "Hsp" designates that the protein is a heat shock protein.

Hsp60[edit]

Hsp60 (GroEL/GroES complex in E. coli) is the best characterized large (~ 1 MDa) chaperone complex. GroEL is a double-ring 14mer with a hydrophobic patch at its opening; it is so large it can accommodate native folding of 54-kDa GFP in its lumen. GroES is a single-ring heptamer that binds to GroEL in the presence of ATP or ADP. GroEL/GroES may not be able to undo previous aggregation, but it does compete in the pathway of misfolding and aggregation.[24] Also acts in mitochondrial matrix as molecular chaperone.

Hsp70[edit]

Hsp70 (DnaK in E. coli) is perhaps the best characterized small (~ 70 kDa) chaperone.
hsp70 pocket for substrate binding
The Hsp70 proteins are aided by Hsp40 proteins (DnaJ in E. coli), which increase the ATP consumption rate and activity of the Hsp70s.
It has been noted that increased expression of Hsp70 proteins in the cell results in a decreased tendency toward apoptosis.
Although a precise mechanistic understanding has yet to be determined, it is known that Hsp70s have a high-affinity bound state to unfolded proteins when bound to ADP, and a low-affinity state when bound to ATP.
It is thought that many Hsp70s crowd around an unfolded substrate, stabilizing it and preventing aggregation until the unfolded molecule folds properly, at which time the Hsp70s lose affinity for the molecule and diffuse away.[25] Hsp70 also acts as a mitochondrial and chloroplastic molecular chaperone in eukaryotes.

Hsp90[edit]

Hsp90 (HtpG in E. coli) may be the least understood chaperone. Its molecular weight is about 90 kDa, and it is necessary for viability in eukaryotes (possibly for prokaryotes as well).
Heat shock protein 90 (Hsp90) is a molecular chaperone essential for activating many signaling proteins in the eukaryotic cell.
Each Hsp90 has an ATP-binding domain, a middle domain, and a dimerization domain. Originally thought to clamp onto their substrate protein (also known as a client protein) upon binding ATP, the recently published structures by Vaughan et al. and Ali et al. indicate that client proteins may bind externally to both the N-terminal and middle domains of Hsp90.[26][27]
Hsp90 may also require co-chaperones-like immunophilinsSti1, p50 (Cdc37), and Aha1, and also cooperates with the Hsp70 chaperone system.[28][29]

Hsp100[edit]

Hsp100 (Clp family in E. coli) proteins have been studied in vivo and in vitro for their ability to target and unfold tagged and misfolded proteins.
Proteins in the Hsp100/Clp family form large hexameric structures with unfoldase activity in the presence of ATP. These proteins are thought to function as chaperones by processively threading client proteins through a small 20 Å (2 nm) pore, thereby giving each client protein a second chance to fold.
Some of these Hsp100 chaperones, like ClpA and ClpX, associate with the double-ringed tetradecameric serine protease ClpP; instead of catalyzing the refolding of client proteins, these complexes are responsible for the targeted destruction of tagged and misfolded proteins.
Hsp104, the Hsp100 of Saccharomyces cerevisiae, is essential for the propagation of many yeast prions. Deletion of the HSP104 gene results in cells that are unable to propagate certain prions.

History[edit]

The investigation of chaperones has a long history.[30] The term "molecular chaperone" appeared first in the literature in 1978, and was invented by Ron Laskey to describe the ability of a nuclear protein called nucleoplasmin to prevent the aggregation of folded histone proteins with DNA during the assembly of nucleosomes.[31] The term was later extended by R. John Ellis in 1987 to describe proteins that mediated the post-translational assembly of protein complexes.[32] In 1988, it was realised that similar proteins mediated this process in both prokaryotes and eukaryotes.[33] The details of this process were determined in 1989, when the ATP-dependent protein folding was demonstrated in vitro.[34]

Clinical significance[edit]

There are many disorders associated with mutations in genes encoding chaperones (i.e. multisystem proteinopathy) that can affect muscle, bone and/or the central nervous system.[35]

See also[edit]

References[edit]

  1. ^ Richardson RT, Alekseev OM, Grossman G, Widgren EE, Thresher R, Wagner EJ, et al. (July 2006). "Nuclear autoantigenic sperm protein (NASP), a linker histone chaperone that is required for cell proliferation". The Journal of Biological Chemistry281 (30): 21526–34. doi:10.1074/jbc.M603816200PMID 16728391.
  2. ^ Alekseev OM, Richardson RT, Alekseev O, O'Rand MG (May 2009). "Analysis of gene expression profiles in HeLa cells in response to overexpression or siRNA-mediated depletion of NASP"Reproductive Biology and Endocrinology7: 45. doi:10.1186/1477-7827-7-45PMC 2686705PMID 19439102.
  3. ^ Ellis RJ (July 2006). "Molecular chaperones: assisting assembly in addition to folding". Trends in Biochemical Sciences31 (7): 395–401. doi:10.1016/j.tibs.2006.05.001PMID 16716593.
  4. ^ Pauwels K, Van Molle I, Tommassen J, Van Gelder P (May 2007). "Chaperoning Anfinsen: the steric foldases" (PDF)Molecular Microbiology64 (4): 917–22. doi:10.1111/j.1365-2958.2007.05718.xPMID 17501917. Archived from the original (PDF) on 2012-05-23.
  5. ^ [Chaperone Action at the Single-Molecule Level http://pubs.acs.org/doi/abs/10.1021/cr400326k]
  6. ^ Hoffmann JH, Linke K, Graf PC, Lilie H, Jakob U (January 2004). "Identification of a redox-regulated chaperone network"The EMBO Journal23 (1): 160–8. doi:10.1038/sj.emboj.7600016PMC 1271656PMID 14685279.
  7. ^ Nillegoda NB, Kirstein J, Szlachcic A, Berynskyy M, Stank A, Stengel F, et al. (August 2015). "Crucial HSP70 co-chaperone complex unlocks metazoan protein disaggregation"Nature524(7564): 247–51. Bibcode:2015Natur.524..247Ndoi:10.1038/nature14884PMC 4830470PMID 26245380.
  8. ^ Balchin D, Hayer-Hartl M, Hartl FU (July 2016). "In vivo aspects of protein folding and quality control". Science353 (6294): aac4354. doi:10.1126/science.aac4354hdl:11858/00-001M-0000-002B-0856-CPMID 27365453.
  9. ^ Ellis RJ, van der Vies SM (1991). "Molecular chaperones". Annual Review of Biochemistry60: 321–47. doi:10.1146/annurev.bi.60.070191.001541PMID 1679318.
  10. ^ van den Berg B, Wain R, Dobson CM, Ellis RJ (August 2000). "Macromolecular crowding perturbs protein refolding kinetics: implications for folding inside the cell"The EMBO Journal19(15): 3870–5. doi:10.1093/emboj/19.15.3870PMC 306593PMID 10921869.
  11. ^ van den Berg B, Ellis RJ, Dobson CM (December 1999). "Effects of macromolecular crowding on protein folding and aggregation"The EMBO Journal18 (24): 6927–33. doi:10.1093/emboj/18.24.6927PMC 1171756PMID 10601015.
  12. ^ Ellis RJ, Minton AP (May 2006). "Protein aggregation in crowded environments". Biological Chemistry387 (5): 485–97. doi:10.1515/BC.2006.064PMID 16740119.
  13. ^ Martin J, Hartl FU (February 1997). "The effect of macromolecular crowding on chaperonin-mediated protein folding"Proceedings of the National Academy of Sciences of the United States of America94 (4): 1107–12. Bibcode:1997PNAS...94.1107Mdoi:10.1073/pnas.94.4.1107PMC 19752PMID 9037014.
  14. ^ Ellis RJ (2007). Protein misassembly: macromolecular crowding and molecular chaperonesAdv. Exp. Med. Biol. Advances in Experimental Medicine and Biology. 594. New York, N.Y. : Sprinter Science+Business Media, LLC ; Austin, Tex. : Landes Bioscience/Eurekah.com. pp. 1–13doi:10.1007/978-0-387-39975-1_1ISBN 978-0-387-39974-4PMID 17205670.
  15. ^ Zhou J, Xu Z (October 2005). "The structural view of bacterial translocation-specific chaperone SecB: implications for function"(PDF)Molecular Microbiology58 (2): 349–57. doi:10.1111/j.1365-2958.2005.04842.xhdl:2027.42/74325PMID 16194224.
  16. ^ Specht S, Miller SB, Mogk A, Bukau B (14 November 2011). "Hsp42 is required for sequestration of protein aggregates into deposition sites in Saccharomyces cerevisiae"J. Cell Biol195(4): 617–29. doi:10.1083/jcb.201106037PMC 3257523PMID 22065637.
  17. ^ Ojha J, Masilamoni G, Dunlap D, Udoff RA, Cashikar AG (August 2011). "Sequestration of toxic oligomers by HspB1 as a cytoprotective mechanism"Mol. Cell. Biol31 (15): 3146–57. doi:10.1128/MCB.01187-10PMC 3147607PMID 21670152.
  18. ^ Mannini B, Cascella R, Zampagni M, van Waarde-Verhagen M, Meehan S, Roodveldt C, Campioni S, Boninsegna M, Penco A, Relini A, Kampinga HH, Dobson CM, Wilson MR, Cecchi C, Chiti F (31 July 2012). "Molecular mechanisms used by chaperones to reduce the toxicity of aberrant protein oligomers"Proc. Natl. Acad. Sci. USA109 (31): 12479–84. Bibcode:2012PNAS..10912479Mdoi:10.1073/pnas.1117799109PMC 3411936PMID 22802614.
  19. ^ Sadigh-Eteghad S, Majdi A, Talebi M, Mahmoudi J, Babri S (May 2015). "Regulation of nicotinic acetylcholine receptors in Alzheimer׳s disease: a possible role of chaperones". European Journal of Pharmacology755: 34–41. doi:10.1016/j.ejphar.2015.02.047PMID 25771456.
  20. ^ Salamanca HH, Antonyak MA, Cerione RA, Shi H, Lis JT (2014). "Inhibiting heat shock factor 1 in human cancer cells with a potent RNA aptamer"PLOS ONE9 (5): e96330. Bibcode:2014PLoSO...996330Sdoi:10.1371/journal.pone.0096330PMC 4011729PMID 24800749.
  21. ^ Ruoppolo M, Orrù S, Talamo F, Ljung J, Pirneskoski A, Kivirikko KI, et al. (May 2003). "Mutations in domain a' of protein disulfide isomerase affect the folding pathway of bovine pancreatic ribonuclease A"Protein Science12 (5): 939–52. doi:10.1110/ps.0242803PMC 2323865PMID 12717017.
  22. ^ Soluble complexes of target proteins and peptidyl prolyl isomerase ...
  23. ^ Frickel EM, Riek R, Jelesarov I, Helenius A, Wuthrich K, Ellgaard L (February 2002). "TROSY-NMR reveals interaction between ERp57 and the tip of the calreticulin P-domain"Proceedings of the National Academy of Sciences of the United States of America99 (4): 1954–9. Bibcode:2002PNAS...99.1954Fdoi:10.1073/pnas.042699099PMC 122301PMID 11842220.
  24. ^ Fenton WA, Horwich AL (May 2003). "Chaperonin-mediated protein folding: fate of substrate polypeptide"Quarterly Reviews of Biophysics36 (2): 229–56. doi:10.1017/S0033583503003883PMID 14686103.
  25. ^ Mayer MP, Bukau B (March 2005). "Hsp70 chaperones: cellular functions and molecular mechanism"Cellular and Molecular Life Sciences62 (6): 670–84. doi:10.1007/s00018-004-4464-6PMC 2773841PMID 15770419.
  26. ^ Vaughan CK, Gohlke U, Sobott F, Good VM, Ali MM, Prodromou C, et al. (September 2006). "Structure of an Hsp90-Cdc37-Cdk4 complex"Molecular Cell23 (5): 697–707. doi:10.1016/j.molcel.2006.07.016PMC 5704897PMID 16949366.
  27. ^ Ali MM, Roe SM, Vaughan CK, Meyer P, Panaretou B, Piper PW, et al. (April 2006). "Crystal structure of an Hsp90-nucleotide-p23/Sba1 closed chaperone complex"Nature440 (7087): 1013–7. Bibcode:2006Natur.440.1013Adoi:10.1038/nature04716PMC 5703407PMID 16625188.
  28. ^ Terasawa K, Minami M, Minami Y (April 2005). "Constantly updated knowledge of Hsp90". Journal of Biochemistry137 (4): 443–7. doi:10.1093/jb/mvi056PMID 15858167.
  29. ^ Pearl LH, Prodromou C (2006). "Structure and mechanism of the Hsp90 molecular chaperone machinery". Annual Review of Biochemistry75: 271–94. doi:10.1146/annurev.biochem.75.103004.142738PMID 16756493.
  30. ^ Ellis RJ (September 1996). "Discovery of molecular chaperones"Cell Stress & Chaperones1 (3): 155–60. doi:10.1379/1466-1268(1996)001<0155:DOMC>2.3.CO;2PMC 248474PMID 9222600.
  31. ^ Laskey RA, Honda BM, Mills AD, Finch JT (October 1978). "Nucleosomes are assembled by an acidic protein which binds histones and transfers them to DNA". Nature275 (5679): 416–20. Bibcode:1978Natur.275..416Ldoi:10.1038/275416a0PMID 692721.
  32. ^ Ellis J (1987). "Proteins as molecular chaperones". Nature328(6129): 378–9. doi:10.1038/328378a0PMID 3112578.
  33. ^ Hemmingsen SM, Woolford C, van der Vies SM, Tilly K, Dennis DT, Georgopoulos CP, et al. (May 1988). "Homologous plant and bacterial proteins chaperone oligomeric protein assembly". Nature333 (6171): 330–4. Bibcode:1988Natur.333..330Hdoi:10.1038/333330a0PMID 2897629.
  34. ^ Goloubinoff P, Christeller JT, Gatenby AA, Lorimer GH (1989). "Reconstitution of active dimeric ribulose bisphosphate carboxylase from an unfoleded state depends on two chaperonin proteins and Mg-ATP". Nature342 (6252): 884–9. Bibcode:1989Natur.342..884Gdoi:10.1038/342884a0PMID 10532860.
  35. ^ Taylor JP (August 2015). "Multisystem proteinopathy: intersecting genetics in muscle, bone, and brain degeneration"Neurology85(8): 658–60. doi:10.1212/WNL.0000000000001862PMID 26208960.

Δεν υπάρχουν σχόλια:

Δημοσίευση σχολίου

Αρχειοθήκη ιστολογίου