Thursday, May 7, 2020

Rosetta@home is a distributed computingproject for protein structure prediction on the Berkeley Open Infrastructure for Network Computing (BOINC) platform, run by the Baker laboratory at the University of Washington. Rosetta@home aims to predict protein–protein docking and design new proteins







Rosetta@home is a distributed computingproject for protein structure prediction on the Berkeley Open Infrastructure for Network Computing (BOINC) platform, run by the Baker laboratory at the University of Washington. Rosetta@home aims to predict protein–protein docking and design new proteins with the help of about fifty-five thousand active volunteered computers processing at over 494,953 GigaFLOPS on average as of March 26, 2020.[5] Foldit, a Rosetta@Home videogame, aims to reach these goals with a crowdsourcingapproach. Though much of the project is oriented toward basic research to improve the accuracy and robustness of proteomicsmethods, Rosetta@home also does applied research on malaria, Alzheimer's disease, and other pathologies.[6]

Like all BOINC projects, Rosetta@home uses idle computer processing resources from volunteers' computers to perform calculations on individual workunits. Completed results are sent to a central project server where they are validated and assimilated into project databases. The project is cross-platform, and runs on a wide variety of hardware configurations. Users can view the progress of their individual protein structure prediction on the Rosetta@home screensaver.

In addition to disease-related research, the Rosetta@home network serves as a testing framework for new methods in structural bioinformatics. Such methods are then used in other Rosetta-based applications, like RosettaDockor the Human Proteome Folding Project and the Microbiome Immunity Project, after being sufficiently developed and proven stable on Rosetta@home's large and diverse set of volunteer computers. Two especially important tests for the new methods developed in Rosetta@home are the Critical Assessment of Techniques for Protein Structure Prediction (CASP) and Critical Assessment of Prediction of Interactions (CAPRI) experiments, biennial experiments which evaluate the state of the art in protein structure prediction and protein–protein docking prediction, respectively. Rosetta@home consistently ranks among the foremost docking predictors, and is one of the best tertiary structure predictors available.[7]

With an influx of new users looking to participate in the fight against the 2019–20 coronavirus pandemic, caused by SARS-CoV-2, Rosetta@home has increased its computing power up to 1.7 PetaFlops as of March 28, 2020.[8][9]

Contents
1Computing platform
2Project significance
3Disease-related research
3.1Alzheimer's disease
3.2Anthrax
3.3Herpes simplex virus 1
3.4HIV
3.5Malaria
3.6Other diseases
4Development history and branches
4.1RosettaDesign
4.2RosettaDock
4.3Robetta
4.4Foldit
5Comparison to similar distributed computing projects
5.1Folding@home
5.2World Community Grid
5.3Predictor@home
6Volunteer contributions
7References
8External links
Computing platform[edit]
See also: List of distributed computing projects

The Rosetta@home application and the BOINC distributed computing platform are available for the operating systems Windows, Linux, and macOS; BOINC also runs on several others, e.g., FreeBSD.[10] Participation in Rosetta@home requires a central processing unit (CPU) with a clock speed of at least 500 MHz, 200 megabytes of free disk space, 512 megabytes of physical memory, and Internet connectivity.[11] As of July 20, 2016, the current version of the Rosetta Mini application is 3.73.[12] The current recommended BOINC program version is 7.6.22.[10] Standard Hypertext Transfer Protocol (HTTP) (port 80) is used for communication between the user's BOINC client and the Rosetta@home servers at the University of Washington; HTTPS (port 443) is used during password exchange. Remote and local control of the BOINC client use port 31416 and port 1043, which might need to be specifically unblocked if they are behind a firewall.[13] Workunits containing data on individual proteins are distributed from servers located in the Baker lab at the University of Washington to volunteers' computers, which then calculate a structure prediction for the assigned protein. To avoid duplicate structure predictions on a given protein, each workunit is initialized with a random seed number. This gives each prediction a unique trajectory of descent along the protein's energy landscape.[14] Protein structure predictions from Rosetta@home are approximations of a global minimum in a given protein's energy landscape. That global minimum represents the most energetically favorable conformation of the protein, i.e., its native state.

Rosetta@home screensaver, showing the progress of a structure prediction for a synthetic ubiquitin protein (PDB ID: 1ogw)

A primary feature of the Rosetta@home graphical user interface (GUI) is a screensaverwhich shows a current workunit's progress during the simulated protein foldingprocess. In the upper-left of the current screensaver, the target protein is shown adopting different shapes (conformations) in its search for the lowest energy structure. Depicted immediately to the right is the structure of the most recently accepted. On the upper right the lowest energy conformation of the current decoy is shown; below that is the true, or native, structure of the protein if it has already been determined. Three graphs are included in the screensaver. Near the middle, a graph for the accepted model's thermodynamic free energy is displayed, which fluctuates as the accepted model changes. A graph of the accepted model's root-mean-square deviation (RMSD), which measures how structurally similar the accepted model is to the native model, is shown far right. On the right of the accepted energy graph and below the RMSD graph, the results from these two functions are used to produce an energy vs. RMSD plot as the model is progressively refined.[15]

Like all BOINC projects, Rosetta@home runs in the background of the user's computer, using idle computer power, either at or before logging into an account on the host operating system. The program frees resources from the CPU as they are needed by other applications so that normal computer use is unaffected. Many program settings can be specified via user account preferences, including: the maximum percentage of CPU resources the program can use (to control power consumption or heat production from a computer running at sustained capacity), the times of day during which the program can run, and many more.

Rosetta, the software that runs on the Rosetta@home network, was rewritten in C++ to allow easier development than that allowed by its original version, which was written in Fortran. This new version is object-oriented, and was released on February 8, 2008.[12][16] Development of the Rosetta code is done by Rosetta Commons.[17] The software is freely licensed to the academic community and available to pharmaceutical companies for a fee.[17]
Project significance[edit]
Further information: Protein structure prediction, Protein docking, and Protein design

With the proliferation of genome sequencing projects, scientists can infer the amino acid sequence, or primary structure, of many proteins that carry out functions within the cell. To better understand a protein's function and aid in rational drug design, scientists need to know the protein's three-dimensional tertiary structure.

CASP6 target T0281, the first ab initio protein structure prediction to approach atomic-level resolution. Rosetta produced a model for T0281 (superpositioned in magenta) 1.5 Ångström (Å) RMSD from the crystal structure (blue).

Protein 3D structures are currently determined experimentally via X-ray crystallography or nuclear magnetic resonance (NMR) spectroscopy. The process is slow (it can take weeks or even months to figure out how to crystallize a protein for the first time) and costly (around US$100,000 per protein).[18] Unfortunately, the rate at which new sequences are discovered far exceeds the rate of structure determination – out of more than 7,400,000 protein sequences available in the National Center for Biotechnology Information (NCBI) nonredundant (nr) protein database, fewer than 52,000 proteins' 3D structures have been solved and deposited in the Protein Data Bank, the main repository for structural information on proteins.[19] One of the main goals of Rosetta@home is to predict protein structures with the same accuracy as existing methods, but in a way that requires significantly less time and money. Rosetta@home also develops methods to determine the structure and docking of membrane proteins (e.g., G protein–coupled receptors (GPCRs)),[20] which are exceptionally difficult to analyze with traditional techniques like X-ray crystallography and NMR spectroscopy, yet represent the majority of targets for modern drugs.

Progress in protein structure prediction is evaluated in the biannual Critical Assessment of Techniques for Protein Structure Prediction (CASP) experiment, in which researchers from around the world attempt to derive a protein's structure from the protein's amino acid sequence. High scoring groups in this sometimes competitive experiment are considered the de facto standard-bearers for what is the state of the art in protein structure prediction. Rosetta, the program on which Rosetta@home is based, has been used since CASP5 in 2002. In the 2004 CASP6 experiment, Rosetta made history by being the first to produce a close to atomic-level resolution, ab initio protein structure prediction in its submitted model for CASP target T0281.[21] Ab initio modeling is considered an especially difficult category of protein structure prediction, as it does not use information from structural homology and must rely on information from sequence homology and modeling physical interactions within the protein. Rosetta@home has been used in CASP since 2006, where it was among the top predictors in every category of structure prediction in CASP7.[22][23][24] These high quality predictions were enabled by the computing power made available by Rosetta@home volunteers.[25] Increasing computing power allows Rosetta@home to sample more regions of conformation space (the possible shapes a protein can assume), which, according to Levinthal's paradox, is predicted to increase exponentially with protein length.

Rosetta@home is also used in protein–protein docking prediction, which determines the structure of multiple complexed proteins, or quaternary structure. This type of protein interaction affects many cellular functions, including antigen–antibody and enzyme–inhibitor binding and cellular import and export. Determining these interactions is critical for drug design. Rosetta is used in the Critical Assessment of Prediction of Interactions (CAPRI) experiment, which evaluates the state of the protein docking field similar to how CASP gauges progress in protein structure prediction. The computing power made available by Rosetta@home's project volunteers has been cited as a major factor in Rosetta's performance in CAPRI, where its docking predictions have been among the most accurate and complete.[26]

In early 2008, Rosetta was used to computationally design a protein with a function never before observed in nature.[27] This was inspired in part by the retraction of a high-profile paper from 2004 which originally described the computational design of a protein with improved enzymatic activity relative to its natural form.[28] The 2008 research paper from David Baker's group describing how the protein was made, which cited Rosetta@home for the computing resources it made available, represented an important proof of concept for this protein design method.[27] This type of protein design could have future applications in drug discovery, green chemistry, and bioremediation.[27]
Disease-related research[edit]

In addition to basic research in predicting protein structure, docking and design, Rosetta@home is also used in immediate disease-related research.[29] Numerous minor research projects are described in David Baker's Rosetta@home journal.[30] As of February 2014, information on recent publications and a short description of the work are being updated on the forum.[31] The forum thread is no longer used since 2016, and news on the research can be found on the general news section of the project.[32]
Alzheimer's disease[edit]

A component of the Rosetta software suite, RosettaDesign, was used to accurately predict which regions of amyloidogenic proteins were most likely to make amyloid-like fibrils.[33] By taking hexapeptides (six amino acid-long fragments) of a protein of interest and selecting the lowest energy match to a structure similar to that of a known fibril forming hexapeptide, RosettaDesign was able to identify peptides twice as likely to form fibrils as are random proteins.[34]Rosetta@home was used in the same study to predict structures for amyloid beta, a fibril-forming protein that has been postulated to cause Alzheimer's disease.[35] Preliminary but as yet unpublished results have been produced on Rosetta-designed proteins that may prevent fibrils from forming, although it is unknown whether it can prevent the disease.[36]
Anthrax[edit]

Another component of Rosetta, RosettaDock,[37][38][39] was used in conjunction with experimental methods to model interactions between three proteins—lethal factor (LF), edema factor (EF) and protective antigen (PA)—that make up anthrax toxin. The computer model accurately predicted docking between LF and PA, helping to establish which domains of the respective proteins are involved in the LF–PA complex. This insight was eventually used in research resulting in improved anthrax vaccines.[40][41]
Herpes simplex virus 1[edit]

RosettaDock was used to model docking between an antibody (immunoglobulin G) and a surface protein expressed by the cold sore virus, herpes simplex virus 1 (HSV-1) which serves to degrade the antiviral antibody. The protein complex predicted by RosettaDock closely agreed with the especially difficult-to-obtain experimental models, leading researchers to conclude that the docking method has potential to address some of the problems that X-ray crystallography has with modeling protein–protein interfaces.[42]
HIV[edit]

As part of research funded by a $19.4 million grant by the Bill & Melinda Gates Foundation,[43]Rosetta@home has been used in designing multiple possible vaccines for human immunodeficiency virus (HIV).[44][45]
Malaria[edit]

In research involved with the Grand Challenges in Global Health initiative,[46] Rosetta has been used to computationally design novel homing endonuclease proteins, which could eradicate Anopheles gambiae or otherwise render the mosquito unable to transmit malaria.[47] Being able to model and alter protein–DNA interactions specifically, like those of homing endonucleases, gives computational protein design methods like Rosetta an important role in gene therapy (which includes possible cancer treatments).[29][48]
Other diseases[edit]

Rosetta@home researchers have designed an IL-2 receptor agonist called Neoleukin-2/15 that does not interact with the alpha subunit of the receptor. Such immunity signal molecules are useful in cancer treatment. While the natural IL-2 suffers from toxicity due to an interaction with the alpha subunit, the designed protein is much safer, at least in animal models.[49] Rosetta molecular modeling suite was recently used to accurately predict the atomic-scale structure of the SARS-CoV-2 spike protein weeks before it could be measured in the lab.[50]
Development history and branches[edit]

Originally introduced by the Baker laboratory in 1998 as an ab initio approach to structure prediction,[51] Rosetta has since branched into several development streams and distinct services. The Rosetta platform derives its name from the Rosetta Stone, as it attempts to decipher the structural "meaning" of proteins' amino acid sequences.[52] More than seven years after Rosetta's first appearance, the Rosetta@home project was released (i.e., announced as no longer beta) on October 6, 2005.[12] Many of the graduate students and other researchers involved in Rosetta's initial development have since moved to other universities and research institutions, and subsequently enhanced different parts of the Rosetta project.
RosettaDesign[edit]

Superposition of Rosetta-designed model (red) for Top7 onto its X-raycrystal structure (blue, PDB ID: 1QYS)

RosettaDesign, a computing approach to protein design based on Rosetta, began in 2000 with a study in redesigning the folding pathway of Protein G.[53] In 2002 RosettaDesign was used to design Top7, a 93-amino acid long α/β protein that had an overall fold never before recorded in nature. This new conformation was predicted by Rosetta to within 1.2 Å RMSD of the structure determined by X-ray crystallography, representing an unusually accurate structure prediction.[54] Rosetta and RosettaDesign earned widespread recognition by being the first to design and accurately predict the structure of a novel protein of such length, as reflected by the 2002 paper describing the dual approach prompting two positive letters in the journal Science,[55][56] and being cited by more than 240 other scientific articles.[57] The visible product of that research, Top7, was featured as the RCSB PDB's 'Molecule of the Month' in October 2006;[58] a superposition of the respective cores (residues 60–79) of its predicted and X-ray crystal structures are featured in the Rosetta@home logo.[21]

Brian Kuhlman, a former postdoctoral associate in David Baker's lab and now an associate professor at the University of North Carolina, Chapel Hill,[59] offers RosettaDesign as an online service.[60]
RosettaDock[edit]

RosettaDock was added to the Rosetta software suite during the first CAPRI experiment in 2002 as the Baker laboratory's algorithm for protein–protein docking prediction.[61] In that experiment, RosettaDock made a high-accuracy prediction for the docking between streptococcal pyogenic exotoxin A and a T cell-receptor β-chain, and a medium accuracy prediction for a complex between porcine α-amylase and a camelid antibody. While the RosettaDock method only made two acceptably accurate predictions out of seven possible, this was enough to rank it seventh out of nineteen prediction methods in the first CAPRI assessment.[61]

Development of RosettaDock diverged into two branches for subsequent CAPRI rounds as Jeffrey Gray, who laid the groundwork for RosettaDock while at the University of Washington, continued working on the method in his new position at Johns Hopkins University. Members of the Baker laboratory further developed RosettaDock in Gray's absence. The two versions differed slightly in side-chain modeling, decoy selection and other areas.[39][62] Despite these differences, both the Baker and Gray methods performed well in the second CAPRI assessment, placing fifth and seventh respectively out of 30 predictor groups.[63] Jeffrey Gray's RosettaDock server is available as a free docking prediction service for non-commercial use.[64]

In October 2006, RosettaDock was integrated into Rosetta@home. The method used a fast, crude docking model phase using only the protein backbone. This was followed by a slow full-atom refinement phase in which the orientation of the two interacting proteins relative to each other, and side-chain interactions at the protein–protein interface, were simultaneously optimized to find the lowest energy conformation.[65] The vastly increased computing power afforded by the Rosetta@home network, combined with revised fold-tree representations for backbone flexibility and loop modeling, made RosettaDock sixth out of 63 prediction groups in the third CAPRI assessment.[7][26]
Robetta[edit]

The Robetta (Rosetta Beta) server is an automated protein structure prediction service offered by the Baker laboratory for non-commercial ab initio and comparative modeling.[66] It has participated as an automated prediction server in the biannual CASP experiments since CASP5 in 2002, performing among the best in the automated server prediction category.[67] Robetta has since competed in CASP6 and 7, where it did better than average among both automated server and human predictor groups.[24][68][69] It also participates in the CAMEO3D continuous evaluation.

In modeling protein structure as of CASP6, Robetta first searches for structural homologs using BLAST, PSI-BLAST, and 3D-Jury, then parses the target sequence into its individual domains, or independently folding units of proteins, by matching the sequence to structural families in the Pfam database. Domains with structural homologs then follow a "template-based model" (i.e., homology modeling) protocol. Here, the Baker laboratory's in-house alignment program, K*sync, produces a group of sequence homologs, and each of these is modeled by the Rosetta de novo method to produce a decoy (possible structure). The final structure prediction is selected by taking the lowest energy model as determined by a low-resolution Rosetta energy function. For domains that have no detected structural homologs, a de novo protocol is followed in which the lowest energy model from a set of generated decoys is selected as the final prediction. These domain predictions are then connected together to investigate inter-domain, tertiary-level interactions within the protein. Finally, side-chain contributions are modeled using a protocol for Monte Carlo conformational search.[70]

In CASP8, Robetta was augmented to use Rosetta's high resolution all-atom refinement method,[71]the absence of which was cited as the main cause for Robetta being less accurate than the Rosetta@home network in CASP7.[25] In CASP11, a way to predict the protein contact map by co-evolution of residues in related proteins called GREMLIN was added, allowing for more de novofold successes.[72]
Foldit[edit]
See also: Foldit

On May 9, 2008, after Rosetta@home users suggested an interactive version of the distributed computing program, the Baker lab publicly released Foldit, an online protein structure prediction game based on the Rosetta platform.[73] As of September 25, 2008, Foldit had over 59,000 registered users.[74] The game gives users a set of controls (for example, shake, wiggle, rebuild) to manipulate the backbone and amino acid side chains of the target protein into more energetically favorable conformations. Users can work on solutions individually as soloists or collectively as evolvers, accruing points under either category as they improve their structure predictions.[75]
Comparison to similar distributed computing projects[edit]

There are several distributed computed projects which have study areas similar to those of Rosetta@home, but differ in their research approach:
Folding@home[edit]

Of all the major distributed computing projects involved in protein research, Folding@home is the only one not using the BOINC platform.[76][77][78] Both Rosetta@home and Folding@home study protein misfolding diseases such as Alzheimer's disease, but Folding@home does so much more exclusively.[79][80] Folding@home almost exclusively uses all-atom molecular dynamics models to understand how and why proteins fold (or potentially misfold, and subsequently aggregate to cause diseases).[81][82] In other words, Folding@home's strength is modeling the process of protein folding, while Rosetta@home's strength is computing protein design and predicting protein structure and docking.

Some of Rosetta@home's results are used as the basis for some Folding@home projects. Rosetta provides the most likely structure, but it is not definite if that is the form the molecule takes or whether or not it is viable. Folding@home can then be used to verify Rosetta@home's results, and can provide added atomic-level information, and details of how the molecule changes shape.[82][83]

The two projects also differ significantly in their computing power and host diversity. Averaging about 6,650 teraFLOPS from a host base of central processing units (CPUs), graphics processing units (GPUs), and PS3s,[84] Folding@home has nearly 108 times more computing power than Rosetta@home.[85]
World Community Grid[edit]

Both Phase I and Phase II of the Human Proteome Folding Project (HPF), a subproject of World Community Grid, have used the Rosetta program to make structural and functional annotations of various genomes.[86][87] Although he now uses it to create databases for biologists, Richard Bonneau, head scientist of the Human Proteome Folding Project, was active in the original development of Rosetta at David Baker's laboratory while obtaining his PhD.[88] More information on the relationship between the HPF1, HPF2 and Rosetta@home can be found on Richard Bonneau's website.[89]
Predictor@home[edit]

Like Rosetta@home, Predictor@home specialized in protein structure prediction.[90] While Rosetta@home uses the Rosetta program for its structure prediction, Predictor@home used the dTASSER methodology.[91] In 2009, Predictor@home shut down.

Other protein related distributed computing projects on BOINC include QMC@home, Docking@home, POEM@home, SIMAP, and TANPAKU. RALPH@home, the Rosetta@home alpha project which tests new application versions, work units, and updates before they move on to Rosetta@home, runs on BOINC also.[92]
Volunteer contributions[edit]

Rosetta@home depends on computing power donated by individual project members for its research. As of March 28, 2020, about 53,000 users from 150 countries were active members of Rosetta@home, together contributing idle processor time from about 54,800 computers for a combined average performance of over 1.7 PetaFLOPS.[85][93]

Bar chart showing cumulative credit per day for Rosetta@home over a 60-day period, indicating its computing power during the CASP8 experiment

Users are granted BOINC credits as a measure of their contribution. The credit granted for each workunit is the number of decoys produced for that workunit multiplied by the average claimed credit for the decoys submitted by all computer hosts for that workunit. This custom system was designed to address significant differences between credit granted to users with the standard BOINC client and an optimized BOINC client, and credit differences between users running Rosetta@home on Windows and Linuxoperating systems.[94] The amount of credit granted per second of CPU work is lower for Rosetta@home than most other BOINC projects.[95]Rosetta@home is thirteenth out of over 40 BOINC projects in terms of total credit.[96]

Rosetta@home users who predict protein structures submitted for the CASP experiment are acknowledged in scientific publications regarding their results.[25] Users who predict the lowest energy structure for a given workunit are featured on the Rosetta@home homepage as Predictor of the Day, along with any team of which they are a member.[97] A User of the Day is chosen randomly each day to be on the homepage also, from among users who have made a Rosetta@home profile.[98]
References[edit]

^ "Rosetta@home License Agreement". Boinc.bakerlab.org. Archived from the original on April 19, 2014. Retrieved April 18, 2014.
^ "Portfolio Highlight: Rosetta++ Software Suite". UW TechTransfer – Digital Ventures. Retrieved September 7, 2008.
^ "Rosetta@home".
^ "Rosetta@Home – Detailed stats | BOINCstats/BAM!".
^ "Rosette@home".
^ "What is Rosetta@home?". Rosetta@home forums. University of Washington. Archived from the original on September 13, 2008. Retrieved September 7, 2008.
^ Jump up to: a b Lensink MF, Méndez R, Wodak SJ (December 2007). "Docking and scoring protein complexes: CAPRI 3rd Edition". Proteins. 69 (4): 704–18. doi:10.1002/prot.21804. PMID 17918726.
^ "Rosetta@home - Server Status "TeraFLOPS estimate"". Rosetta@home. March 25, 2020. Retrieved March 25, 2020.
^ "Rosetta@home Rallies a Legion of Computers Against the Coronavirus". HPCWire. March 24, 2020. Retrieved March 25, 2020.
^ Jump up to: a b "Download BOINC client software". BOINC. University of California. 2008. Retrieved December 1, 2008.
^ "Rosetta@home: Recommended System Requirements". Rosetta@home. University of Washington. 2008. Archived from the original on September 25, 2008. Retrieved October 7, 2008.
^ Jump up to: a b c "Rosetta@home: News archive". Rosetta@home. University of Washington. 2016. Retrieved July 20, 2016.
^ "Rosetta@home: FAQ (work in progress) (message 10910)". Rosetta@home forums. University of Washington. 2006. Retrieved October 7, 2008.
^ Kim DE (2005). "Rosetta@home: Random Seed (message 3155)". Rosetta@home forums. University of Washington. Retrieved October 7, 2008.
^ "Rosetta@home: Quick guide to Rosetta and its graphics". Rosetta@home. University of Washington. 2007. Archived from the original on September 24, 2008. Retrieved October 7, 2008.
^ Kim DE (2008). "Rosetta@home: Problems with minirosetta version 1.+ (Message 51199)". Rosetta@home forums. University of Washington. Retrieved September 7, 2008.
^ Jump up to: a b "Rosetta Commons". RosettaCommons.org. 2008. Archived from the original on September 15, 2008. Retrieved October 7, 2008.
^ Bourne PE, Helge W, eds. (2003). Structural Bioinformatics. Hoboken, NJ: Wiley-Liss. ISBN 978-0-471-20199-1. OCLC 50199108.
^ "Yearly Growth of Protein Structures". RCSB Protein Data Bank. 2008. Retrieved November 30,2008.
^ Baker D (2008). "Rosetta@home: David Baker's Rosetta@home journal (message 55893)". Rosetta@home forums. University of Washington. Retrieved October 7, 2008.
^ Jump up to: a b "Rosetta@home: Research Overview". Rosetta@home. University of Washington. 2007. Archived from the original on September 25, 2008. Retrieved October 7, 2008.
^ Kopp J, Bordoli L, Battey JN, Kiefer F, Schwede T (2007). "Assessment of CASP7 predictions for template-based modeling targets". Proteins. 69 Suppl 8: 38–56. doi:10.1002/prot.21753. PMID 17894352.
^ Read RJ, Chavali G (2007). "Assessment of CASP7 predictions in the high accuracy template-based modeling category". Proteins. 69 Suppl 8: 27–37. doi:10.1002/prot.21662. PMID 17894351.
^ Jump up to: a b Jauch R, Yeo HC, Kolatkar PR, Clarke ND (2007). "Assessment of CASP7 structure predictions for template free targets". Proteins. 69 Suppl 8: 57–67. doi:10.1002/prot.21771. PMID 17894330.
^ Jump up to: a b c Das R, Qian B, Raman S, et al. (2007). "Structure prediction for CASP7 targets using extensive all-atom refinement with Rosetta@home". Proteins. 69 Suppl 8: 118–28. doi:10.1002/prot.21636. PMID 17894356.
^ Jump up to: a b Wang C, Schueler-Furman O, Andre I, et al. (December 2007). "RosettaDock in CAPRI rounds 6–12". Proteins. 69 (4): 758–63. doi:10.1002/prot.21684. PMID 17671979.
^ Jump up to: a b c Jiang L, Althoff EA, Clemente FR, et al. (March 2008). "De novo computational design of retro-aldol enzymes". Science. 319 (5868): 1387–91. Bibcode:2008Sci...319.1387J. doi:10.1126/science.1152692. PMC 3431203. PMID 18323453.
^ Hayden EC (February 13, 2008). "Protein prize up for grabs after retraction". Nature. doi:10.1038/news.2008.569.
^ Jump up to: a b "Disease Related Research". Rosetta@home. University of Washington. 2008. Archived from the original on September 23, 2008. Retrieved October 8, 2008.
^ Baker D (2008). "Rosetta@home: David Baker's Rosetta@home journal". Rosetta@home forums. University of Washington. Retrieved September 7, 2008.
^ "Rosetta@home Research Updates". Boinc.bakerlab.org. Retrieved April 18, 2014.
^ "News archive". Rosetta@home. Retrieved May 10, 2019.
^ Kuhlman B, Baker D (September 2000). "Native protein sequences are close to optimal for their structures". Proceedings of the National Academy of Sciences of the United States of America. 97(19): 10383–88. Bibcode:2000PNAS...9710383K. doi:10.1073/pnas.97.19.10383. PMC 27033. PMID 10984534.
^ Thompson MJ, Sievers SA, Karanicolas J, Ivanova MI, Baker D, Eisenberg D (March 2006). "The 3D profile method for identifying fibril-forming segments of proteins". Proceedings of the National Academy of Sciences of the United States of America. 103 (11): 4074–78. Bibcode:2006PNAS..103.4074T. doi:10.1073/pnas.0511295103. PMC 1449648. PMID 16537487.
^ Bradley P. "Rosetta@home forum: Amyloid fibril structure prediction". Rosetta@home forums. University of Washington. Retrieved September 7, 2008.
^ Baker D. "Rosetta@home forum: Publications on R@H's Alzheimer's work? (message 54681)". Rosetta@home forums. University of Washington. Retrieved October 8, 2008.
^ Wang C, Schueler-Furman O, Baker D (May 2005). "Improved side-chain modeling for protein–protein docking". Protein Science. 14 (5): 1328–39. doi:10.1110/ps.041222905. PMC 2253276. PMID 15802647.
^ Gray JJ, Moughon S, Wang C, et al. (August 2003). "Protein–protein docking with simultaneous optimization of rigid-body displacement and side-chain conformations". Journal of Molecular Biology. 331 (1): 281–99. doi:10.1016/S0022-2836(03)00670-3. PMID 12875852.
^ Jump up to: a b Schueler-Furman O, Wang C, Baker D (August 2005). "Progress in protein–protein docking: atomic resolution predictions in the CAPRI experiment using RosettaDock with an improved treatment of side-chain flexibility". Proteins. 60 (2): 187–94. doi:10.1002/prot.20556. PMID 15981249.
^ Lacy DB, Lin HC, Melnyk RA, et al. (November 2005). "A model of anthrax toxin lethal factor bound to protective antigen". Proceedings of the National Academy of Sciences of the United States of America. 102 (45): 16409–14. Bibcode:2005PNAS..10216409L. doi:10.1073/pnas.0508259102. PMC 1283467. PMID 16251269.
^ Albrecht MT, Li H, Williamson ED, et al. (November 2007). "Human monoclonal antibodies against anthrax lethal factor and protective antigen act independently to protect against Bacillus anthracis infection and enhance endogenous immunity to anthrax". Infection and Immunity. 75 (11): 5425–33. doi:10.1128/IAI.00261-07. PMC 2168292. PMID 17646360.
^ Sprague ER, Wang C, Baker D, Bjorkman PJ (June 2006). "Crystal structure of the HSV-1 Fc receptor bound to Fc reveals a mechanism for antibody bipolar bridging". PLOS Biology. 4 (6): e148. doi:10.1371/journal.pbio.0040148. PMC 1450327. PMID 16646632.
^ Paulson, Tom (July 19, 2006). "Gates Foundation awards $287 million for HIV vaccine research". Seattle Post-Intelligencer. Retrieved September 7, 2008.
^ Liu Y, et al. (2007). "Development of IgG1 b12 scaffolds and HIV-1 env-based outer domain immunogens capable of eliciting and detecting IgG1 b12-like antibodies" (PDF). Global HIV Vaccine Enterprise. Archived from the original (PDF) on February 25, 2009. Retrieved September 28, 2008.
^ Baker D. "David Baker's Rosetta@home journal archives (message 40756)". Rosetta@home forums. University of Washington. Retrieved September 7, 2008.
^ "Homing Endonuclease Genes: New Tools for Mosquito Population Engineering and Control". Grand Challenges in Global Health. Retrieved September 7, 2008.
^ Windbichler N, Papathanos PA, Catteruccia F, Ranson H, Burt A, Crisanti A (2007). "Homing endonuclease mediated gene targeting in Anopheles gambiae cells and embryos". Nucleic Acids Research. 35 (17): 5922–33. doi:10.1093/nar/gkm632. PMC 2034484. PMID 17726053.
^ Ashworth J, Havranek JJ, Duarte CM, et al. (June 2006). "Computational redesign of endonuclease DNA binding and cleavage specificity". Nature. 441 (7093): 656–59. Bibcode:2006Natur.441..656A. doi:10.1038/nature04818. PMC 2999987. PMID 16738662.
^ Silva DA, Yu S, Ulge UY, Spangler JB, Jude KM, Labão-Almeida C, Ali LR, Quijano-Rubio A, Ruterbusch M, Leung I, Biary T, Crowley SJ, Marcos E, Walkey CD, Weitzner BD, Pardo-Avila F, Castellanos J, Carter L, Stewart L, Riddell SR, Pepper M, Bernardes GJ, Dougan M, Garcia KC, Baker D (January 2019). "De novo design of potent and selective mimics of IL-2 and IL-15". Nature. 565 (7738): 186–191. doi:10.1038/s41586-018-0830-7. PMC 6521699. PMID 30626941.
^ "Rosetta's role in fighting coronavirus – Institute for Protein Design". Retrieved March 6, 2020.
^ Simons KT, Bonneau R, Ruczinski I, Baker D (1999). "Ab initio protein structure prediction of CASP III targets using Rosetta". Proteins. Suppl 3: 171–76. doi:10.1002/(SICI)1097-0134(1999)37:3+<171::AID-PROT21>3.0.CO;2-Z. PMID 10526365.
^ "Interview with David Baker". Team Picard Distributed Computing. 2006. Archived from the original on February 18, 2009. Retrieved December 23, 2008.
^ Nauli S, Kuhlman B, Baker D (July 2001). "Computer-based redesign of a protein folding pathway". Nature Structural Biology. 8 (7): 602–05. doi:10.1038/89638. PMID 11427890.
^ Kuhlman B, Dantas G, Ireton GC, Varani G, Stoddard BL, Baker D (November 2003). "Design of a novel globular protein fold with atomic-level accuracy". Science. 302 (5649): 1364–68. Bibcode:2003Sci...302.1364K. doi:10.1126/science.1089427. PMID 14631033.
^ Jones DT (November 2003). "Structural biology. Learning to speak the language of proteins". Science. 302 (5649): 1347–48. doi:10.1126/science.1092492. PMID 14631028.
^ von Grotthuss M, Wyrwicz LS, Pas J, Rychlewski L (June 2004). "Predicting protein structures accurately". Science. 304 (5677): 1597–99, author reply 1597–99. doi:10.1126/science.304.5677.1597b. PMID 15192202.
^ "Articles citing: Kuhlman et al. (2003) 'Design of a novel globular protein fold with atomic-level accuracy'". ISI Web of Science. Retrieved July 10, 2008.
^ "October 2005 molecule of the month: Designer proteins". RCSB Protein Data Bank. Retrieved September 7, 2008.
^ "Kuhlman laboratory homepage". Kuhlman Laboratory. University of North Carolina. Retrieved September 7, 2008.
^ "RosettaDesign web server". Kuhlman Laboratory. University of North Carolina. Retrieved September 7, 2008.
^ Jump up to: a b Gray JJ, Moughon SE, Kortemme T, et al. (July 2003). "Protein–protein docking predictions for the CAPRI experiment". Proteins. 52 (1): 118–22. CiteSeerX 10.1.1.80.9354. doi:10.1002/prot.10384. PMID 12784377.
^ Daily MD, Masica D, Sivasubramanian A, Somarouthu S, Gray JJ (2005). "CAPRI rounds 3–5 reveal promising successes and future challenges for RosettaDock". Proteins. 60 (2): 181–86. CiteSeerX 10.1.1.521.9981. doi:10.1002/prot.20555. PMID 15981262. Archived from the original on June 30, 2012.
^ Méndez R, Leplae R, Lensink MF, Wodak SJ (2005). "Assessment of CAPRI predictions in rounds 3–5 shows progress in docking procedures". Proteins. 60 (2): 150–69. doi:10.1002/prot.20551. PMID 15981261. Archived from the original on June 30, 2012.
^ "RosettaDock server". Rosetta Commons. Retrieved March 28, 2020.
^ "Protein–protein docking at Rosetta@home". Rosetta@home forums. University of Washington. Retrieved September 7, 2008.
^ "Robetta web server". Baker laboratory. University of Washington. Retrieved May 7, 2019.
^ Aloy P, Stark A, Hadley C, Russell RB (2003). "Predictions without templates: new folds, secondary structure, and contacts in CASP5". Proteins. 53 Suppl 6: 436–56. doi:10.1002/prot.10546. PMID 14579333.
^ Tress M, Ezkurdia I, Graña O, López G, Valencia A (2005). "Assessment of predictions submitted for the CASP6 comparative modeling category". Proteins. 61 Suppl 7: 27–45. doi:10.1002/prot.20720. PMID 16187345.
^ Battey JN, Kopp J, Bordoli L, Read RJ, Clarke ND, Schwede T (2007). "Automated server predictions in CASP7". Proteins. 69 Suppl 8: 68–82. doi:10.1002/prot.21761. PMID 17894354.
^ Chivian D, Kim DE, Malmström L, Schonbrun J, Rohl CA, Baker D (2005). "Prediction of CASP6 structures using automated Robetta protocols". Proteins. 61 Suppl 7: 157–66. doi:10.1002/prot.20733. PMID 16187358.
^ Baker D. "David Baker's Rosetta@home journal, message 52902". Rosetta@home forums. University of Washington. Retrieved September 7, 2008.
^ Ovchinnikov, S; Kim, DE; Wang, RY; Liu, Y; DiMaio, F; Baker, D (September 2016). "Improved de novo structure prediction in CASP11 by incorporating coevolution information into Rosetta". Proteins. 84 Suppl 1: 67–75. doi:10.1002/prot.24974. PMC 5490371. PMID 26677056.
^ Baker D. "David Baker's Rosetta@home journal (message 52963)". Rosetta@home forums. University of Washington. Retrieved September 16, 2008.
^ "Foldit forums: How many users does Foldit have? Etc. (message 2)". University of Washington. Retrieved September 27, 2008.
^ "Foldit: Frequently Asked Questions". fold.it. University of Washington. Retrieved September 19,2008.
^ "Project list – BOINC". University of California. Retrieved September 8, 2008.
^ Pande Group (2010). "High Performance FAQ". Stanford University. Archived from the original (FAQ) on September 21, 2012. Retrieved September 19, 2011.
^ 7im (April 2, 2010). "Re: Answers to: Reasons for not using F@H". Retrieved September 19,2011.
^ Vijay Pande (August 5, 2011). "Results page updated – new key result published in our work in Alzheimer's Disease". Retrieved September 19, 2011.
^ Pande Group. "Folding@home Diseases Studied FAQ". Stanford University. Archived from the original (FAQ) on October 11, 2007. Retrieved September 12, 2011.
^ Vijay Pande (September 26, 2007). "How FAH works: Molecular dynamics". Retrieved September 10, 2011.
^ Jump up to: a b tjlane (June 9, 2011). "Re: Course grained Protein folding in under 10 minutes". Retrieved September 19, 2011.
^ jmn (July 29, 2011). "Rosetta@home and Folding@home: additional projects". Retrieved September 19, 2011.
^ Pande Group. "Client Statistics by OS". Stanford University. Retrieved October 18, 2011.
^ Jump up to: a b "Rosetta@home: Credit overview". boincstats.com. Retrieved March 28, 2020.
^ Malmström L, Riffle M, Strauss CE, et al. (April 2007). "Superfamily assignments for the yeast proteome through integration of structure prediction with the gene ontology". PLOS Biology. 5 (4): e76. doi:10.1371/journal.pbio.0050076. PMC 1828141. PMID 17373854.
^ Bonneau R (2006). "World Community Grid Message Board Posts: HPF -> HPF2 transition". Bonneau Lab, New York University. Retrieved September 7, 2008.
^ "List of Richard Bonneau's publications". Bonneau Lab, New York University. Archived from the original on July 7, 2008. Retrieved September 7, 2008.
^ Bonneau R. "World Community Grid Message Board Posts". Bonneau Lab, New York University. Archived from the original on July 4, 2008. Retrieved September 7, 2008.
^ "Predictor@home: Developing new application areas for P@H". The Brooks Research Group. Retrieved September 7, 2008.[dead link]
^ Carrillo-Tripp M (2007). "dTASSER". The Scripps Research Institute. Archived from the original on July 6, 2007. Retrieved September 7, 2008.
^ "RALPH@home website". RALPH@home forums. University of Washington. Retrieved September 7, 2008.
^ "Rosetta@home". Retrieved March 19, 2020.
^ "Rosetta@home: The new credit system explained". Rosetta@home forums. University of Washington. 2006. Retrieved October 8, 2008.
^ "BOINCstats: Project Credit Comparison". boincstats.com. 2008. Archived from the originalon September 13, 2008. Retrieved October 8, 2008.
^ "Credit divided over projects". boincstats.com. Retrieved February 19, 2015.
^ "Rosetta@home: Predictor of the day archive". Rosetta@home. University of Washington. 2008. Archived from the original on September 24, 2008. Retrieved October 8, 2008.
^ "Rosetta@home: Protein Folding, Design, and Docking". Rosetta@home. University of Washington. 2008. Retrieved October 8, 2008.
External links[edit]
Official website
Baker Lab Baker Lab website
David Baker's Rosetta@home journal
BOINC Includes platform overview, and a guide to install BOINC and attach to Rosetta@home
BOINCstats – Rosetta@home Detailed contribution statistics
RALPH@home Website for Rosetta@home alpha testing project
Rosetta@home video on YouTube Overview of Rosetta@home given by David Baker and lab members
Rosetta Commons Academic collaborative for developing the Rosetta platform
Kuhlman lab webpage, home of RosettaDesign

Online Rosetta services
Rosetta Commons list of available servers
Robetta Protein structure prediction server
ROSIE Docking, design, etc. multifunctional server-set
RosettaDesign Protein design server
RosettaBackrub Flexible backbone / protein design server

Folding@home (FAH or F@h) is a distributed computing projecd





Folding@home (FAH or F@h) is a distributed computing project aimed to help scientists develop new therapeutics to a variety of diseases by the means of simulating protein dynamics. This includes the process of protein folding and the movements of proteins, and it's all reliant on the simulations run on the volunteers' personal computers.[4] Folding@home is currently based at Washington University in St. Louis and led by Dr. Greg Bowman, a former student of Dr. Pande.[5]

The project has pioneered the utilization of central processing units (CPUs), graphics processing units(GPUs), PlayStation 3s, Message Passing Interface(used for computing on multi-core processors), and some Sony Xperia smartphones for distributed computing and scientific research. The project uses statistical simulation methodology that is a paradigm shift from traditional computing methods.[6] As part of the client–server model network architecture, the volunteered machines each receive pieces of a simulation (work units), complete them, and return them to the project's database servers, where the units are compiled into an overall simulation. Volunteers can track their contributions on the Folding@home website, which makes volunteers' participation competitive and encourages long-term involvement.

Folding@home is one of the world's fastest computing systems. With heightened interest in the project as a result of the 2019–20 coronavirus pandemic, the system achieved a speed of approximately 1.22 exaFLOPS by late March 2020 and reaching 2.43 x86 exaFLOPS by April 12, 2020,[7] making it the world's first exaFLOP computing system. This level of performance from its large-scale computing network has allowed researchers to run computationally costly atomic-level simulations of protein folding thousands of times longer than formerly achieved. Since its launch on 1 October 2000, the Pande Lab has produced 223 scientific research papers as a direct result of Folding@home.[8] Results from the project's simulations agree well with experiments.[9][10][11]

Contents
1Background
2Examples of application in biomedical research
2.1Alzheimer's disease
2.2Huntington's disease
2.3Cancer
2.4Osteogenesis imperfecta
2.5Viruses
2.6Drug design
3Potential applications in biomedical research
3.1Prion diseases
4Patterns of participation
4.1Performance
4.2Points
5Software
5.1Work units
5.2Cores
5.3Client
5.3.1Graphics processing units
5.3.2PlayStation 3
5.3.3Multi-core processing client
5.3.4V7
5.3.5Google Chrome
5.3.6Android
6Comparison to other molecular simulators
7See also
8References
9Sources
10External links
Background[edit]
Further information: Protein folding

A protein before and after folding. It starts in an unstable random coil state and finishes in its native state conformation.

Proteins are an essential component to many biological functions and participate in virtually all processes within biological cells. They often act as enzymes, performing biochemical reactions including cell signaling, molecular transportation, and cellular regulation. As structural elements, some proteins act as a type of skeleton for cells, and as antibodies, while other proteins participate in the immune system. Before a protein can take on these roles, it must fold into a functional three-dimensional structure, a process that often occurs spontaneously and is dependent on interactions within its amino acid sequence and interactions of the amino acids with their surroundings. Protein folding is driven by the search to find the most energetically favorable conformation of the protein, i.e., its native state. Thus, understanding protein folding is critical to understanding what a protein does and how it works, and is considered a holy grail of computational biology.[12][13] Despite folding occurring within a crowded cellular environment, it typically proceeds smoothly. However, due to a protein's chemical properties or other factors, proteins may misfold, that is, fold down the wrong pathway and end up misshapen. Unless cellular mechanisms can destroy or refold misfolded proteins, they can subsequently aggregate and cause a variety of debilitating diseases.[14] Laboratory experiments studying these processes can be limited in scope and atomic detail, leading scientists to use physics-based computing models that, when complementing experiments, seek to provide a more complete picture of protein folding, misfolding, and aggregation.[15][16]

Due to the complexity of proteins' conformation or configuration space (the set of possible shapes a protein can take), and limits in computing power, all-atom molecular dynamics simulations have been severely limited in the timescales which they can study. While most proteins typically fold in the order of milliseconds,[15][17] before 2010, simulations could only reach nanosecond to microsecond timescales.[9] General-purpose supercomputers have been used to simulate protein folding, but such systems are intrinsically costly and typically shared among many research groups. Further, because the computations in kinetic models occur serially, strong scaling of traditional molecular simulations to these architectures is exceptionally difficult.[18][19] Moreover, as protein folding is a stochastic process and can statistically vary over time, it is challenging computationally to use long simulations for comprehensive views of the folding process.[20][21]

Folding@home uses Markov state models, like the one diagrammed here, to model the possible shapes and folding pathways a protein can take as it condenses from its initial randomly coiled state (left) into its native 3-D structure (right).

Protein folding does not occur in one step.[14] Instead, proteins spend most of their folding time, nearly 96% in some cases,[22] waiting in various intermediate conformational states, each a local thermodynamic free energy minimum in the protein's energy landscape. Through a process known as adaptive sampling, these conformations are used by Folding@home as starting points for a set of simulation trajectories. As the simulations discover more conformations, the trajectories are restarted from them, and a Markov state model (MSM) is gradually created from this cyclic process. MSMs are discrete-time master equation models which describe a biomolecule's conformational and energy landscape as a set of distinct structures and the short transitions between them. The adaptive sampling Markov state model method significantly increases the efficiency of simulation as it avoids computation inside the local energy minimum itself, and is amenable to distributed computing (including on GPUGRID) as it allows for the statistical aggregation of short, independent simulation trajectories.[23] The amount of time it takes to construct a Markov state model is inversely proportional to the number of parallel simulations run, i.e., the number of processors available. In other words, it achieves linear parallelization, leading to an approximately four orders of magnitude reduction in overall serial calculation time. A completed MSM may contain tens of thousands of sample states from the protein's phase space (all the conformations a protein can take on) and the transitions between them. The model illustrates folding events and pathways (i.e., routes) and researchers can later use kinetic clustering to view a coarse-grained representation of the otherwise highly detailed model. They can use these MSMs to reveal how proteins misfold and to quantitatively compare simulations with experiments.[6][20][24]

Between 2000 and 2010, the length of the proteins Folding@home has studied have increased by a factor of four, while its timescales for protein folding simulations have increased by six orders of magnitude.[25] In 2002, Folding@home used Markov state models to complete approximately a million CPU days of simulations over the span of several months,[11] and in 2011, MSMs parallelized another simulation that required an aggregate 10 million CPU hours of computing.[26] In January 2010, Folding@home used MSMs to simulate the dynamics of the slow-folding 32-residueNTL9 protein out to 1.52 milliseconds, a timescale consistent with experimental folding rate predictions but a thousand times longer than formerly achieved. The model consisted of many individual trajectories, each two orders of magnitude shorter, and provided an unprecedented level of detail into the protein's energy landscape.[6][9][27] In 2010, Folding@home researcher Gregory Bowman was awarded the Thomas Kuhn Paradigm Shift Award from the American Chemical Society for the development of the open-source MSMBuilder software and for attaining quantitative agreement between theory and experiment.[28][29] For his work, Pande was awarded the 2012 Michael and Kate Bárány Award for Young Investigators for "developing field-defining and field-changing computational methods to produce leading theoretical models for protein and RNAfolding",[30] and the 2006 Irving Sigal Young Investigator Award for his simulation results which "have stimulated a re-examination of the meaning of both ensemble and single-molecule measurements, making Dr. Pande's efforts pioneering contributions to simulation methodology."[31]
Examples of application in biomedical research[edit]

Protein misfolding can result in a variety of diseases including Alzheimer's disease, cancer, Creutzfeldt–Jakob disease, cystic fibrosis, Huntington's disease, sickle-cell anemia, and type II diabetes.[14][32][33] Cellular infection by viruses such as HIV and influenza also involve folding events on cell membranes.[34] Once protein misfolding is better understood, therapies can be developed that augment cells' natural ability to regulate protein folding. Such therapies include the use of engineered molecules to alter the production of a given protein, help destroy a misfolded protein, or assist in the folding process.[35] The combination of computational molecular modeling and experimental analysis has the possibility to fundamentally shape the future of molecular medicine and the rational design of therapeutics,[16] such as expediting and lowering the costs of drug discovery.[36] The goal of the first five years of Folding@home was to make advances in understanding folding, while the current goal is to understand misfolding and related disease, especially Alzheimer's.[37]

The simulations run on Folding@home are used in conjunction with laboratory experiments,[20] but researchers can use them to study how folding in vitro differs from folding in native cellular environments. This is advantageous in studying aspects of folding, misfolding, and their relationships to disease that are difficult to observe experimentally. For example, in 2011, Folding@home simulated protein folding inside a ribosomal exit tunnel, to help scientists better understand how natural confinement and crowding might influence the folding process.[38][39]Furthermore, scientists typically employ chemical denaturants to unfold proteins from their stable native state. It is not generally known how the denaturant affects the protein's refolding, and it is difficult to experimentally determine if these denatured states contain residual structures which may influence folding behavior. In 2010, Folding@home used GPUs to simulate the unfolded states of Protein L, and predicted its collapse rate in strong agreement with experimental results.[40]

The large data sets from the project are freely available for other researchers to use upon request and some can be accessed from the Folding@home website.[41][42] The Pande lab has collaborated with other molecular dynamics systems such as the Blue Gene supercomputer,[43] and they share Folding@home's key software with other researchers, so that the algorithms which benefited Folding@home may aid other scientific areas.[41] In 2011, they released the open-source Copernicus software, which is based on Folding@home's MSM and other parallelizing methods and aims to improve the efficiency and scaling of molecular simulations on large computer clustersor supercomputers.[44][45] Summaries of all scientific findings from Folding@home are posted on the Folding@home website after publication.[46]
Alzheimer's disease[edit]



Alzheimer's disease is linked to the aggregation of amyloid beta protein fragments in the brain (right). Researchers have used Folding@home to simulate this aggregation process, to better understand the cause of the disease.

Alzheimer's disease is an incurable neurodegenerative disease which most often affects the elderly and accounts for more than half of all cases of dementia. Its exact cause remains unknown, but the disease is identified as a protein misfolding disease. Alzheimer's is associated with toxic aggregations of the amyloid beta (Aβ) peptide, caused by Aβ misfolding and clumping together with other Aβ peptides. These Aβ aggregates then grow into significantly larger senile plaques, a pathological marker of Alzheimer's disease.[47][48][49] Due to the heterogeneous nature of these aggregates, experimental methods such as X-ray crystallography and nuclear magnetic resonance(NMR) have had difficulty characterizing their structures. Moreover, atomic simulations of Aβ aggregation are highly demanding computationally due to their size and complexity.[50][51]

Preventing Aβ aggregation is a promising method to developing therapeutic drugs for Alzheimer's disease, according to Drs. Naeem and Fazili in a literature review article.[52] In 2008, Folding@home simulated the dynamics of Aβ aggregation in atomic detail over timescales of the order of tens of seconds. Prior studies were only able to simulate about 10 microseconds. Folding@home was able to simulate Aβ folding for six orders of magnitude longer than formerly possible. Researchers used the results of this study to identify a beta hairpin that was a major source of molecular interactions within the structure.[53] The study helped prepare the Pande lab for future aggregation studies and for further research to find a small peptide which may stabilize the aggregation process.[50]

In December 2008, Folding@home found several small drug candidates which appear to inhibit the toxicity of Aβ aggregates.[54] In 2010, in close cooperation with the Center for Protein Folding Machinery, these drug leads began to be tested on biological tissue.[33] In 2011, Folding@home completed simulations of several mutations of Aβ that appear to stabilize the aggregate formation, which could aid in the development of therapeutic drug therapies for the disease and greatly assist with experimental nuclear magnetic resonance spectroscopy studies of Aβ oligomers.[51][55] Later that year, Folding@home began simulations of various Aβ fragments to determine how various natural enzymes affect the structure and folding of Aβ.[56][57]
Huntington's disease[edit]

Huntington's disease is a neurodegenerative genetic disorder that is associated with protein misfolding and aggregation. Excessive repeats of the glutamine amino acid at the N-terminus of the huntingtin protein cause aggregation, and although the behavior of the repeats is not completely understood, it does lead to the cognitive decline associated with the disease.[58] As with other aggregates, there is difficulty in experimentally determining its structure.[59] Scientists are using Folding@home to study the structure of the huntingtin protein aggregate and to predict how it forms, assisting with rational drug design methods to stop the aggregate formation.[33] The N17 fragment of the huntingtin protein accelerates this aggregation, and while there have been several mechanisms proposed, its exact role in this process remains largely unknown.[60] Folding@home has simulated this and other fragments to clarify their roles in the disease.[61] Since 2008, its drug design methods for Alzheimer's disease have been applied to Huntington's.[33]
Cancer[edit]

More than half of all known cancers involve mutations of p53, a tumor suppressor protein present in every cell which regulates the cell cycle and signals for cell death in the event of damage to DNA. Specific mutations in p53 can disrupt these functions, allowing an abnormal cell to continue growing unchecked, resulting in the development of tumors. Analysis of these mutations helps explain the root causes of p53-related cancers.[62] In 2004, Folding@home was used to perform the first molecular dynamics study of the refolding of p53's protein dimer in an all-atom simulation of water. The simulation's results agreed with experimental observations and gave insights into the refolding of the dimer that were formerly unobtainable.[63] This was the first peer reviewed publication on cancer from a distributed computing project.[64] The following year, Folding@home powered a new method to identify the amino acids crucial for the stability of a given protein, which was then used to study mutations of p53. The method was reasonably successful in identifying cancer-promoting mutations and determined the effects of specific mutations which could not otherwise be measured experimentally.[65]

Folding@home is also used to study protein chaperones,[33] heat shock proteins which play essential roles in cell survival by assisting with the folding of other proteins in the crowded and chemically stressful environment within a cell. Rapidly growing cancer cells rely on specific chaperones, and some chaperones play key roles in chemotherapy resistance. Inhibitions to these specific chaperones are seen as potential modes of action for efficient chemotherapy drugs or for reducing the spread of cancer.[66] Using Folding@home and working closely with the Center for Protein Folding Machinery, the Pande lab hopes to find a drug which inhibits those chaperones involved in cancerous cells.[67] Researchers are also using Folding@home to study other molecules related to cancer, such as the enzyme Src kinase, and some forms of the engrailedhomeodomain: a large protein which may be involved in many diseases, including cancer.[68][69] In 2011, Folding@home began simulations of the dynamics of the small knottin protein EETI, which can identify carcinomas in imaging scans by binding to surface receptors of cancer cells.[70][71]

Interleukin 2 (IL-2) is a protein that helps T cells of the immune system attack pathogens and tumors. However, its use as a cancer treatment is restricted due to serious side effects such as pulmonary edema. IL-2 binds to these pulmonary cells differently than it does to T cells, so IL-2 research involves understanding the differences between these binding mechanisms. In 2012, Folding@home assisted with the discovery of a mutant form of IL-2 which is three hundred times more effective in its immune system role but carries fewer side effects. In experiments, this altered form significantly outperformed natural IL-2 in impeding tumor growth. Pharmaceutical companieshave expressed interest in the mutant molecule, and the National Institutes of Health are testing it against a large variety of tumor models to try to accelerate its development as a therapeutic.[72][73]
Osteogenesis imperfecta[edit]

Osteogenesis imperfecta, known as brittle bone disease, is an incurable genetic bone disorder which can be lethal. Those with the disease are unable to make functional connective bone tissue. This is most commonly due to a mutation in Type-I collagen,[74] which fulfills a variety of structural roles and is the most abundant protein in mammals.[75] The mutation causes a deformation in collagen's triple helix structure, which if not naturally destroyed, leads to abnormal and weakened bone tissue.[76] In 2005, Folding@home tested a new quantum mechanical method that improved upon prior simulation methods, and which may be useful for future computing studies of collagen.[77] Although researchers have used Folding@home to study collagen folding and misfolding, the interest stands as a pilot project compared to Alzheimer's and Huntington's research.[33]
Viruses[edit]

Folding@home is assisting in research towards preventing some viruses, such as influenza and HIV, from recognizing and entering biological cells.[33] In 2011, Folding@home began simulations of the dynamics of the enzyme RNase H, a key component of HIV, to try to design drugs to deactivate it.[78] Folding@home has also been used to study membrane fusion, an essential event for viral infection and a wide range of biological functions. This fusion involves conformational changes of viral fusion proteins and protein docking,[34] but the exact molecular mechanisms behind fusion remain largely unknown.[79] Fusion events may consist of over a half million atoms interacting for hundreds of microseconds. This complexity limits typical computer simulations to about ten thousand atoms over tens of nanoseconds: a difference of several orders of magnitude.[53] The development of models to predict the mechanisms of membrane fusion will assist in the scientific understanding of how to target the process with antiviral drugs.[80] In 2006, scientists applied Markov state models and the Folding@home network to discover two pathways for fusion and gain other mechanistic insights.[53]

Following detailed simulations from Folding@home of small cells known as vesicles, in 2007, the Pande lab introduced a new computing method to measure the topology of its structural changes during fusion.[81] In 2009, researchers used Folding@home to study mutations of influenza hemagglutinin, a protein that attaches a virus to its host cell and assists with viral entry. Mutations to hemagglutinin affect how well the protein binds to a host's cell surface receptor molecules, which determines how infective the virus strain is to the host organism. Knowledge of the effects of hemagglutinin mutations assists in the development of antiviral drugs.[82][83] As of 2012, Folding@home continues to simulate the folding and interactions of hemagglutinin, complementing experimental studies at the University of Virginia.[33][84]

In March 2020, Folding@home launched a program to assist researchers around the world who are working on finding a cure and learning more about the coronavirus pandemic. The initial wave of projects simulate potentially druggable protein targets from SARS-CoV-2 virus, and the related SARS-CoV virus, about which there is significantly more data available.[85][86][87]
Drug design[edit]

Drugs function by binding to specific locations on target molecules and causing some desired change, such as disabling a target or causing a conformational change. Ideally, a drug should act very specifically, and bind only to its target without interfering with other biological functions. However, it is difficult to precisely determine where and how tightly two molecules will bind. Due to limits in computing power, current in silico methods usually must trade speed for accuracy; e.g., use rapid protein docking methods instead of computationally costly free energy calculations. Folding@home's computing performance allows researchers to use both methods, and evaluate their efficiency and reliability.[37][88][89] Computer-assisted drug design has the potential to expedite and lower the costs of drug discovery.[36] In 2010, Folding@home used MSMs and free energy calculations to predict the native state of the villin protein to within 1.8 angstrom (Å) root mean square deviation (RMSD) from the crystalline structure experimentally determined through X-ray crystallography. This accuracy has implications to future protein structure prediction methods, including for intrinsically unstructured proteins.[53] Scientists have used Folding@home to research drug resistance by studying vancomycin, an antibiotic drug of last resort, and beta-lactamase, a protein that can break down antibiotics like penicillin.[90][91]

Chemical activity occurs along a protein's active site. Traditional drug design methods involve tightly binding to this site and blocking its activity, under the assumption that the target protein exists in one rigid structure. However, this approach works for approximately only 15% of all proteins. Proteins contain allosteric sites which, when bound to by small molecules, can alter a protein's conformation and ultimately affect the protein's activity. These sites are attractive drug targets, but locating them is very computationally costly. In 2012, Folding@home and MSMs were used to identify allosteric sites in three medically relevant proteins: beta-lactamase, interleukin-2, and RNase H.[91][92]

Approximately half of all known antibiotics interfere with the workings of a bacteria's ribosome, a large and complex biochemical machine that performs protein biosynthesis by translatingmessenger RNA into proteins. Macrolide antibiotics clog the ribosome's exit tunnel, preventing synthesis of essential bacterial proteins. In 2007, the Pande lab received a grant to study and design new antibiotics.[33] In 2008, they used Folding@home to study the interior of this tunnel and how specific molecules may affect it.[93] The full structure of the ribosome was determined only as of 2011, and Folding@home has also simulated ribosomal proteins, as many of their functions remain largely unknown.[94]
Potential applications in biomedical research[edit]

There are many more protein misfolding promoted diseases that can be benefited from Folding@home to either discern the misfolded protein structure or the misfolding kinetics, and assist in drug design in the future. The often fatal prion diseases is among the most significant.
Prion diseases[edit]


This section possibly contains synthesis of material which does not verifiably mention or relate to the main topic.Relevant discussion may be found on the talk page. (March 2020)(Learn how and when to remove this template message)


Prion (PrP) is a transmembrane cellular protein found widely in eukaryotic cells. In mammals, it is more abundant in the central nervous system. Although its function is unknown, its high conservation among species indicates an important role in the cellular function. The conformational change from the normal prion protein (PrPc, stands for cellular) to the disease causing isoformPrPSc (stands for prototypical prion disease–scrapie) causes a host of diseases collectly known as transmissible spongiform encephalopathies (TSEs), including Bovine spongiform encephalopathy(BSE) in bovine, Creutzfeldt-Jakob disease (CJD) and fatal insomnia in human, chronic wasting disease (CWD) in the deer family. The conformational change is widely accepted as the result of protein misfolding. What distinguishes TSEs from other protein misfolding diseases is its transmissible nature. The ‘seeding’ of the infectious PrPSc, either arising spontaneously, hereditary or acquired via exposure to contaminated tissues,[95] can cause a chain reaction of transforming normal PrPc into fibrils aggregates or amyloid like plaques consist of PrPSc.[96]

The molecular structure of PrPSc has not been fully characterized due to its aggregated nature. Neither is known much about the mechanism of the protein misfolding nor its kinetics. Using the known structure of PrPc and the results of the in vitro and in vivo studies described below, Folding@home could be valuable in elucidating how PrPSc is formed and how the infectious protein arrange themselves to form fibrils and amyloid like plaques, bypassing the requirement to purify PrPSc or dissolve the aggregates.

The PrPc has been enzymatically dissociated from the membrane and purified, its structure studied using structure characterization techniques such as NMR spectroscopy and X-ray crystallography. Post-translational PrPc has 231 amino acids (aa) in murine. The molecule consists of a long and unstructured amino terminal region spanning up to aa residue 121 and a structured carboxy terminal domain.[96] This globular domain harbours two short sheet-forming anti-parallel β-strands(aa 128 to 130 and aa 160 to 162 in murine PrPc) and three α-helices (helix I: aa 143 to 153; helix II: aa 171 to 192; helix III: aa 199 to 226 in murine PrPc),[97] Helices II and III are anti-parallel orientated and connected by a short loop. Their structural stability is supported by a disulfide bridge, which is parallel to both sheet-forming β-strands. These α-helices and the β-sheet form the rigid core of the globular domain of PrPc.[98]

The disease causing PrPSc is proteinase K resistant and insoluble. Attempts to purify it from the brains of infected animals invariably yield heterogeneous mixtures and aggregated states that are not amenable to characterization by NMR spectroscopy or X-ray crystallography. However, it is a general consensus that PrPSc contains a high percentage of tightly stacked β-sheets than the normal PrPc that renders the protein insoluble and resistant to proteinase. Using techniques of cryoelectron microscopy and structural modeling based on similar common protein structures, it has been discovered that PrPSc contains ß-sheets in the region of aa 81-95 to aa 171, while the carboxy terminal structure is supposedly preserved, retaining the disulfide-linked α-helical conformation in the normal PrPc. These ß-sheets form a parallel left-handed beta-helix.[96] Three PrPSc molecules are believed to form a primary unit and therefore build the basis for the so-called scrapie-associated fibrils.[99] The catalytic activity depends on the size of the particle. PrPSc particles which consist of only 14-28 PrPc molecules exhibit the highest rate of infectivity and conversion.[100]

Despite the difficulty to purify and characterize PrPSc, from the known molecular structure of PrPc and using transgenic mice and N-terminal deletion,[101] the potential ‘hot spots’ of protein misfolding leading to the pathogenic PrPSc could be deduced and Folding@home could be of great value in confirming these. Studies found that both the primary and secondary structure of the prion protein can be of significance of the conversion.

There are more than twenty mutations of the prion protein gene (PRNP) that are known to be associated with or that are directly linked to the hereditary form of human TSEs [56], indicating single amino acids at certain position, likely within the carboxy domain,[97] of the PrPc can affect the susceptibility to TSEs.

The post-translational amino terminal region of PrPc consists of residues 23-120 which make up nearly half of the amino sequence of full-length matured PrPc. There are two sections in the amino terminal region that may influence conversion. First, residues 52-90 contains an octapeptide repeat (5 times) region that likely influences the initial binding (via the octapeptide repeats) and also the actual conversion via the second section of aa 108–124.[102] The highly hydrophobic AGAAAAGA is located between aa residue 113 and 120 and is described as putative aggregation site,[103]although this sequence requires its flanking parts to form fibrillar aggregates.[104]

In the carboxy globular domain,[98] among the three helices, study show that helix II has a significant higher propensity to β-strand conformation.[105] Due to the high conformational flexvoribility seen between residues 114-125 (part of the unstructured N-terminus chain) and the high β-strand propensity of helix II, only moderate changes in the environmental conditions or interactions might be sufficient to induce misfolding of PrPc and subsequent fibril formation.[96]

Other studies of NMR structures of PrPc showed that these residues (~108–189) contain most of the folded domain including both β-strands, the first two α-helices, and the loop/turn regions connecting them, but not the helix III.[101] Small changes within the loop/turn structures of PrPc itself could be important in the conversion as well.[106] In another study, Riek et al. showed that the two small regions of β-strand upstream of the loop regions act as a nucleation site for the conformational conversion of the loop/turn and α-helical structures in PrPc to β-sheet.[97]

The energy threshold for the conversion are not necessarily high. The folding stability, i.e. the free energy of a globular protein in its environment is in the range of one or two hydrogen bonds thus allows the transition to an isoform without the requirement of high transition energy.[96]

From the respective of the interactions among the PrPc molecules, hydrophobic interactions play a crucial role in the formation of β-sheets, a hallmark of PrPSc, as the sheets bring fragments of polypeptide chains into close proximity.[107] Indeed, Kutznetsov and Rackovsky [108] showed that disease-promoting mutations in the human PrPc had a statistically significant tendency towards increasing local hydrophobicity.

In vitro experiments showed the kinetics of misfolding has an initial lag phase followed by a rapid growth phase of fibril formation.[109] It is likely that PrPc goes through some intermediate states, such as at least partially unfolded or degraded, before finally ending up as part of an amyloid fibril.[96]
Patterns of participation[edit]

Like other distributed computing projects, Folding@home is an online citizen science project. In these projects non-specialists contribute computer processing power or help to analyse data produced by professional scientists. Participants in these projects play an invaluable role in facilitating research for little or no obvious reward.

Research has been carried out into the motivations of citizen scientists and most of these studies have found that participants are motivated to take part because of altruistic reasons, that is, they want to help scientists and make a contribution to the advancement of their research.[110][111][112][113] Many participants in citizen science have an underlying interest in the topic of the research and gravitate towards projects that are in disciplines of interest to them. Folding@home is no different in that respect.[114] Research carried out recently on over 400 active participants revealed that they wanted to help make a contribution to research and that many had friends or relatives affected by the diseases that the Folding@home scientists investigate.

Folding@home attracts participants who are computer hardware enthusiasts (sometimes called ‘overclockers’). These groups bring considerable expertise to the project and are able to build computers with advanced processing power.[115] Other distributed computing projects attract these types of participants and projects are often used to benchmark the performance of modified computers, and this aspect of the hobby is accommodated through the competitive nature of the project. Individuals and teams can compete to see who can process the most computer processing units (CPUs).

This latest research on Folding@home involving interview and ethnographic observation of online groups showed that teams of hardware enthusiasts can sometimes work together, sharing best practice with regard to maximising processing output. Such teams can become communities of practice, with a shared language and online culture. This pattern of participation has been observed in other distributed computing projects.[116][117]

Another key observation of Folding@home participants is that many are male.[114] This has also been observed in other distributed projects. Furthermore, many participants work in computer and technology-based jobs and careers.[114][118][119]

Not all Folding@home participants are hardware enthusiasts. Many participants run the project software on unmodified machines and do take part competitively. Over 100,000 participants are involved in Folding@home. However, it is difficult to ascertain what proportion of participants are hardware enthusiasts. Although, according to the project managers, the contribution of the enthusiast community is substantially larger in terms of processing power.[120]
Performance[edit]

Computing power of Folding@home and the fastest supercomputer from April 2004 to October 2012. Between June 2007 and June 2011, Folding@home (red) exceeded the performance of Top500's fastest supercomputer (black). However it was eclipsed by K computer in November 2011 and Blue Gene/Q in June 2012.

Supercomputer FLOPS performance is assessed by running the legacy LINPACK benchmark. This short-term testing has difficulty in accurately reflecting sustained performance on real-world tasks because LINPACK more efficiently maps to supercomputer hardware. Computing systems vary in architecture and design, so direct comparison is difficult. Despite this, FLOPS remain the primary speed metric used in supercomputing.[121][need quotation to verify] In contrast, Folding@home determines its FLOPS using wall-clock time by measuring how much time its work units take to complete.[122]

On September 16, 2007, due in large part to the participation of PlayStation 3 consoles, the Folding@home project officially attained a sustained performance level higher than one native petaFLOPS, becoming the first computing system of any kind to do so.[123][124] Top500's fastest supercomputer at the time was BlueGene/L, at 0.280 petaFLOPS.[125] The following year, on May 7, 2008, the project attained a sustained performance level higher than two native petaFLOPS,[126] followed by the three and four native petaFLOPS milestones on August 2008[127][128] and September 28, 2008 respectively.[129] On February 18, 2009, Folding@home achieved five native petaFLOPS,[130][131] and was the first computing project to meet these five levels.[132][133] In comparison, November 2008's fastest supercomputer was IBM's Roadrunner at 1.105 petaFLOPS.[134] On November 10, 2011, Folding@home's performance exceeded six native petaFLOPS with the equivalent of nearly eight x86 petaFLOPS.[124][135] In mid-May 2013, Folding@home attained over seven native petaFLOPS, with the equivalent of 14.87 x86 petaFLOPS. It then reached eight native petaFLOPS on June 21, followed by nine on September 9 of that year, with 17.9 x86 petaFLOPS.[136] On May 11, 2016 Folding@home announced that it was moving towards reaching the 100 x86 petaFLOPS mark.[137]

Further use grew from increased awareness and participation in the project from the coronavirus pandemic in 2020. On March 20, 2020 Folding@home announced via Twitter that it was running with over 470 native petaFLOPS[138], the equivalent of 958 x86 petaFLOPS.[139] By March 25 it reached 768 petaFLOPS, or 1.5 x86 exaFLOPS, making it the first exaFLOP computing system.[140]
Points[edit]

Similarly to other distributed computing projects, Folding@home quantitatively assesses user computing contributions to the project through a credit system.[141] All units from a given protein project have uniform base credit, which is determined by benchmarking one or more work units from that project on an official reference machine before the project is released.[141] Each user receives these base points for completing every work unit, though through the use of a passkey they can receive added bonus points for reliably and rapidly completing units which are more demanding computationally or have a greater scientific priority.[142][143] Users may also receive credit for their work by clients on multiple machines.[144] This point system attempts to align awarded credit with the value of the scientific results.[141]

Users can register their contributions under a team, which combine the points of all their members. A user can start their own team, or they can join an existing team. In some cases, a team may have their own community-driven sources of help or recruitment such as an Internet forum.[145] The points can foster friendly competition between individuals and teams to compute the most for the project, which can benefit the folding community and accelerate scientific research.[141][146][147]Individual and team statistics are posted on the Folding@home website.[141]

If a user does not form a new team, or does not join an existing team, that user automatically becomes part of a "Default" team. This "Default" team has a team number of "0". Statistics are accumulated for this "Default" team as well as for specially named teams.
Software[edit]

Folding@home software at the user's end involves three primary components: work units, cores, and a client.
Work units[edit]

A work unit is the protein data that the client is asked to process. Work units are a fraction of the simulation between the states in a Markov model. After the work unit has been downloaded and completely processed by a volunteer's computer, it is returned to Folding@home servers, which then award the volunteer the credit points. This cycle repeats automatically.[146] All work units have associated deadlines, and if this deadline is exceeded, the user may not get credit and the unit will be automatically reissued to another participant. As protein folding occurs serially, and many work units are generated from their predecessors, this allows the overall simulation process to proceed normally if a work unit is not returned after a reasonable period of time. Due to these deadlines, the minimum system requirement for Folding@home is a Pentium 3 450 MHz CPU with Streaming SIMD Extensions (SSE).[144] However, work units for high-performance clients have a much shorter deadline than those for the uniprocessor client, as a major part of the scientific benefit is dependent on rapidly completing simulations.[148]

Before public release, work units go through several quality assurance steps to keep problematic ones from becoming fully available. These testing stages include internal, beta, and advanced, before a final full release across Folding@home.[149] Folding@home's work units are normally processed only once, except in the rare event that errors occur during processing. If this occurs for three different users, the unit is automatically pulled from distribution.[150][151] The Folding@home support forum can be used to differentiate between issues arising from problematic hardware and bad work units.[152]
Cores[edit]
Main article: List of Folding@home cores

Specialized molecular dynamics programs, referred to as "FahCores" and often abbreviated "cores", perform the calculations on the work unit as a background process. A large majority of Folding@home's cores are based on GROMACS,[146] one of the fastest and most popular molecular dynamics software packages, which largely consists of manually optimized assembly language code and hardware optimizations.[153][154] Although GROMACS is open-source softwareand there is a cooperative effort between the Pande lab and GROMACS developers, Folding@home uses a closed-source license to help ensure data validity.[155] Less active cores include ProtoMol and SHARPEN. Folding@home has used AMBER, CPMD, Desmond, and TINKER, but these have since been retired and are no longer in active service.[3][156][157] Some of these cores perform explicit solvation calculations in which the surrounding solvent (usually water) is modeled atom-by-atom; while others perform implicit solvation methods, where the solvent is treated as a mathematical continuum.[158][159] The core is separate from the client to enable the scientific methods to be updated automatically without requiring a client update. The cores periodically create calculation checkpoints so that if they are interrupted they can resume work from that point upon startup.[146]
Client[edit]

Folding@Home running on Fedora 25

A Folding@home participant installs a client program on their personal computer. The user interacts with the client, which manages the other software components in the background. Through the client, the user may pause the folding process, open an event log, check the work progress, or view personal statistics.[160] The computer clients run continuously in the background at a very low priority, using idle processing power so that normal computer use is unaffected.[144] The maximum CPU use can be adjusted via client settings.[160][161] The client connects to a Folding@home server and retrieves a work unit and may also download the appropriate core for the client's settings, operating system, and the underlying hardware architecture. After processing, the work unit is returned to the Folding@home servers. Computer clients are tailored to uniprocessor and multi-core processorsystems, and graphics processing units. The diversity and power of each hardware architectureprovides Folding@home with the ability to efficiently complete many types of simulations in a timely manner (in a few weeks or months rather than years), which is of significant scientific value. Together, these clients allow researchers to study biomedical questions formerly considered impractical to tackle computationally.[37][146][148]

Professional software developers are responsible for most of Folding@home's code, both for the client and server-side. The development team includes programmers from Nvidia, ATI, Sony, and Cauldron Development.[162] Clients can be downloaded only from the official Folding@home website or its commercial partners, and will only interact with Folding@home computer files. They will upload and download data with Folding@home's data servers (over port 8080, with 80 as an alternate), and the communication is verified using 2048-bit digital signatures.[144][163] While the client's graphical user interface (GUI) is open-source,[164] the client is proprietary software citing security and scientific integrity as the reasons.[165][166][167]

However, this rationale of using proprietary software is disputed since while the license could be enforceable in the legal domain retrospectively, it doesn't practically prevent the modification (also known as patching) of the executable binary files. Likewise, binary-only distribution does not prevent the malicious modification of executable binary-code, either through a man-in-the-middle attack while being downloaded via the internet,[168] or by the redistribution of binaries by a third-party that have been previously modified either in their binary state (i.e. patched),[169] or by decompiling[170] and recompiling them after modification.[171][172] These modifications are possible unless the binary files – and the transport channel – are signed and the recipient person/system is able to verify the digital signature, in which case unwarranted modifications should be detectable, but not always.[173] Either way, since in the case of Folding@home the input data and output result processed by the client-software are both digitally signed,[144][163] the integrity of work can be verified independently from the integrity of the client software itself.

Folding@home uses the Cosm software libraries for networking.[146][162] Folding@home was launched on October 1, 2000, and was the first distributed computing project aimed at bio-molecular systems.[174] Its first client was a screensaver, which would run while the computer was not otherwise in use.[175][176] In 2004, the Pande lab collaborated with David P. Anderson to test a supplemental client on the open-source BOINC framework. This client was released to closed beta in April 2005;[177] however, the method became unworkable and was shelved in June 2006.[178]
Graphics processing units[edit]

The specialized hardware of graphics processing units (GPU) is designed to accelerate rendering of 3-D graphics applications such as video games and can significantly outperform CPUs for some types of calculations. GPUs are one of the most powerful and rapidly growing computing platforms, and many scientists and researchers are pursuing general-purpose computing on graphics processing units (GPGPU). However, GPU hardware is difficult to use for non-graphics tasks and usually requires significant algorithm restructuring and an advanced understanding of the underlying architecture.[179] Such customization is challenging, more so to researchers with limited software development resources. Folding@home uses the open-source OpenMM library, which uses a bridge design pattern with two application programming interface (API) levels to interface molecular simulation software to an underlying hardware architecture. With the addition of hardware optimizations, OpenMM-based GPU simulations need no significant modification but achieve performance nearly equal to hand-tuned GPU code, and greatly outperform CPU implementations.[158][180]

Before 2010, the computing reliability of GPGPU consumer-grade hardware was largely unknown, and circumstantial evidence related to the lack of built-in error detection and correction in GPU memory raised reliability concerns. In the first large-scale test of GPU scientific accuracy, a 2010 study of over 20,000 hosts on the Folding@home network detected soft errors in the memory subsystems of two-thirds of the tested GPUs. These errors strongly correlated to board architecture, though the study concluded that reliable GPU computing was very feasible as long as attention is paid to the hardware traits, such as software-side error detection.[181]

The first generation of Folding@home's GPU client (GPU1) was released to the public on October 2, 2006,[178] delivering a 20–30 times speedup for some calculations over its CPU-based GROMACS counterparts.[182] It was the first time GPUs had been used for either distributed computing or major molecular dynamics calculations.[183][184] GPU1 gave researchers significant knowledge and experience with the development of GPGPU software, but in response to scientific inaccuracies with DirectX, on April 10, 2008 it was succeeded by GPU2, the second generation of the client.[182][185] Following the introduction of GPU2, GPU1 was officially retired on June 6.[182]Compared to GPU1, GPU2 was more scientifically reliable and productive, ran on ATI and CUDA-enabled Nvidia GPUs, and supported more advanced algorithms, larger proteins, and real-time visualization of the protein simulation.[186][187] Following this, the third generation of Folding@home's GPU client (GPU3) was released on May 25, 2010. While backward compatiblewith GPU2, GPU3 was more stable, efficient, and flexibile in its scientific abilities,[188] and used OpenMM on top of an OpenCL framework.[188][189] Although these GPU3 clients did not natively support the operating systems Linux and macOS, Linux users with Nvidia graphics cards were able to run them through the Wine software application.[190][191] GPUs remain Folding@home's most powerful platform in FLOPS. As of November 2012, GPU clients account for 87% of the entire project's x86 FLOPS throughput.[192]

Native support for Nvidia and AMD graphics cards under Linux was introduced with FahCore 17, which uses OpenCL rather than CUDA.[193]
PlayStation 3[edit]
Further information: Life with PlayStation

The PlayStation 3's Life With PlayStation client displays a 3-D animation of the protein being folded

From March 2007 until November 2012, Folding@home took advantage of the computing power of PlayStation 3s. At the time of its inception, its main streaming Cell processor delivered a 20 times speed increase over PCs for some calculations, processing power which could not be found on other systems such as the Xbox 360.[37][194]The PS3's high speed and efficiency introduced other opportunities for worthwhile optimizations according to Amdahl's law, and significantly changed the tradeoff between computing efficiency and overall accuracy, allowing the use of more complex molecular models at little added computing cost.[195] This allowed Folding@home to run biomedical calculations that would have been otherwise infeasible computationally.[196]

The PS3 client was developed in a collaborative effort between Sony and the Pande lab and was first released as a standalone client on March 23, 2007.[37][197] Its release made Folding@home the first distributed computing project to use PS3s.[198] On September 18 of the following year, the PS3 client became a channel of Life with PlayStation on its launch.[199][200] In the types of calculations it can perform, at the time of its introduction, the client fit in between a CPU's flexibility and a GPU's speed.[146] However, unlike clients running on personal computers, users were unable to perform other activities on their PS3 while running Folding@home.[196] The PS3's uniform console environment made technical support easier and made Folding@home more user friendly.[37] The PS3 also had the ability to stream data quickly to its GPU, which was used for real-time atomic-level visualizing of the current protein dynamics.[195]

On November 6, 2012, Sony ended support for the Folding@home PS3 client and other services available under Life with PlayStation. Over its lifetime of five years and seven months, more than 15 million users contributed over 100 million hours of computing to Folding@home, greatly assisting the project with disease research. Following discussions with the Pande lab, Sony decided to terminate the application. Pande considered the PlayStation 3 client a "game changer" for the project.[201][202][203]
Multi-core processing client[edit]

Folding@home can use the parallel computing abilities of modern multi-core processors. The ability to use several CPU cores simultaneously allows completing the full simulation far faster. Working together, these CPU cores complete single work units proportionately faster than the standard uniprocessor client. This method is scientifically valuable because it enables much longer simulation trajectories to be performed in the same amount of time, and reduces the traditional difficulties of scaling a large simulation to many separate processors.[204] A 2007 publication in the Journal of Molecular Biology relied on multi-core processing to simulate the folding of part of the villin protein approximately 10 times longer than was possible with a single-processor client, in agreement with experimental folding rates.[205]

In November 2006, first-generation symmetric multiprocessing (SMP) clients were publicly released for open beta testing, referred to as SMP1.[178] These clients used Message Passing Interface(MPI) communication protocols for parallel processing, as at that time the GROMACS cores were not designed to be used with multiple threads.[148] This was the first time a distributed computing project had used MPI.[206] Although the clients performed well in Unix-based operating systems such as Linux and macOS, they were troublesome under Windows.[204][206] On January 24, 2010, SMP2, the second generation of the SMP clients and the successor to SMP1, was released as an open beta and replaced the complex MPI with a more reliable thread-based implementation.[143][162]

SMP2 supports a trial of a special category of bigadv work units, designed to simulate proteins that are unusually large and computationally intensive and have a great scientific priority. These units originally required a minimum of eight CPU cores,[207] which was raised to sixteen later, on February 7, 2012.[208] Along with these added hardware requirements over standard SMP2 work units, they require more system resources such as random-access memory (RAM) and Internet bandwidth. In return, users who run these are rewarded with a 20% increase over SMP2's bonus point system.[209] The bigadv category allows Folding@home to run especially demanding simulations for long times that had formerly required use of supercomputing clusters and could not be performed anywhere else on Folding@home.[207] Many users with hardware able to run bigadv units have later had their hardware setup deemed ineligible for bigadv work units when CPU core minimums were increased, leaving them only able to run the normal SMP work units. This frustrated many users who invested significant amounts of money into the program only to have their hardware be obsolete for bigadv purposes shortly after. As a result, Pande announced in January 2014 that the bigadv program would end on January 31, 2015.[210]
V7[edit]

A sample image of the V7 client in Novice mode running under Windows 7. In addition to a variety of controls and user details, V7 presents work unit information, such as its state, calculation progress, ETA, credit points, identification numbers, and description.

The V7 client is the seventh and latest generation of the Folding@home client software, and is a full rewrite and unification of the prior clients for Windows, macOS, and Linux operating systems.[211][212] It was released on March 22, 2012.[213] Like its predecessors, V7 can run Folding@home in the background at a very low priority, allowing other applications to use CPU resources as they need. It is designed to make the installation, start-up, and operation more user-friendly for novices, and offer greater scientific flexibility to researchers than prior clients.[214] V7 uses Trac for managing its bug tickets so that users can see its development process and provide feedback.[212]

V7 consists of four integrated elements. The user typically interacts with V7's open-source GUI, named FAHControl.[164][215] This has Novice, Advanced, and Expert user interface modes, and has the ability to monitor, configure, and control many remote folding clients from one computer. FAHControl directs FAHClient, a back-end application that in turn manages each FAHSlot (or slot). Each slot acts as replacement for the formerly distinct Folding@home v6 uniprocessor, SMP, or GPU computer clients, as it can download, process, and upload work units independently. The FAHViewer function, modeled after the PS3's viewer, displays a real-time 3-D rendering, if available, of the protein currently being processed.[211][212]
Google Chrome[edit]

In 2014, a client for the Google Chrome and Chromium web browsers was released, allowing users to run Folding@home in their web browser. The client used Google's Native Client (NaCl) feature on Chromium-based web browsers to run the Folding@home code at near-native speed in a sandbox on the user's machine.[216] Due to the phasing out of NaCL and changes at Folding@home, the web client was permanently shut down in June 2019.[217]
Android[edit]

In July 2015, a client for Android mobile phones was released on Google Play for devices running Android 4.4 KitKat or newer.[218][219]

On the 16th of February 2018 the Android client, which was offered in cooperation with Sony, was removed from the Google Play. Plans were announced to offer an open source alternative in the future.[220]
Comparison to other molecular simulators[edit]

Rosetta@home is a distributed computing project aimed at protein structure prediction and is one of the most accurate tertiary structure predictors.[221][222] The conformational states from Rosetta's software can be used to initialize a Markov state model as starting points for Folding@home simulations.[23] Conversely, structure prediction algorithms can be improved from thermodynamic and kinetic models and the sampling aspects of protein folding simulations.[223] As Rosetta only tries to predict the final folded state, and not how folding proceeds, Rosetta@home and Folding@home are complementary and address very different molecular questions.[23][224]

Anton is a special-purpose supercomputer built for molecular dynamics simulations. In October 2011, Anton and Folding@home were the two most powerful molecular dynamics systems.[225]Anton is unique in its ability to produce single ultra-long computationally costly molecular trajectories,[226] such as one in 2010 which reached the millisecond range.[227][228] These long trajectories may be especially helpful for some types of biochemical problems.[229][230] However, Anton does not use Markov state models (MSM) for analysis. In 2011, the Pande lab constructed a MSM from two 100-µs Anton simulations and found alternative folding pathways that were not visible through Anton's traditional analysis. They concluded that there was little difference between MSMs constructed from a limited number of long trajectories or one assembled from many shorter trajectories.[226] In June 2011 Folding@home added sampling of an Anton simulation in an effort to better determine how its methods compare to Anton's.[231][232] However, unlike Folding@home's shorter trajectories, which are more amenable to distributed computing and other parallelizing methods, longer trajectories do not require adaptive sampling to sufficiently sample the protein's phase space. Due to this, it is possible that a combination of Anton's and Folding@home's simulation methods would provide a more thorough sampling of this space.[226]
See also[edit]

Biology portal
Medicine portal
BOINC
Foldit
List of distributed computing projects
Comparison of software for molecular mechanics modeling
Molecular modeling on GPUs
SETI@home
Storage@home
Molecule editor
World Community Grid
References[edit]

^ foldingathome.org (September 27, 2016). "About Folding@home Partners".
^
Bowman, Greg (April 17, 2020). "New Folding@home software with the option to prioritize COVID-19 projects". Folding@home. Retrieved April 18, 2020.
"Alternative Downloads". Folding@home. Retrieved April 18, 2020.
^ Jump up to: a b Pande lab (August 2, 2012). "Folding@home Open Source FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on March 3, 2020. Retrieved July 8, 2013.
^ Folding@home n.d.e: "Folding@home (FAH or F@h) is a distributed computing project for simulating protein dynamics, including the process of protein folding and the movements of proteins implicated in a variety of diseases. It brings together citizen scientists who volunteer to run simulations of protein dynamics on their personal computers. Insights from this data are helping scientists to better understand biology, and providing new opportunities for developing therapeutics."
^ Julia Evangelou Strait (February 26, 2019). "Computational biology project aims to better understand protein folding". Retrieved March 8, 2020.
^ Jump up to: a b c V. S. Pande; K. Beauchamp; G. R. Bowman (2010). "Everything you wanted to know about Markov State Models but were afraid to ask". Methods. 52 (1): 99–105. doi:10.1016/j.ymeth.2010.06.002. PMC 2933958. PMID 20570730.
^ Pande lab. "Client Statistics by OS". Archive.is. Retrieved April 12, 2020.
^ "Papers & Results". Folding@home.org. Retrieved March 13, 2020.
^ Jump up to: a b c Vincent A. Voelz; Gregory R. Bowman; Kyle Beauchamp; Vijay S. Pande (2010). "Molecular simulation of ab initio protein folding for a millisecond folder NTL9(1–39)". Journal of the American Chemical Society. 132 (5): 1526–1528. doi:10.1021/ja9090353. PMC 2835335. PMID 20070076.
^ Gregory R. Bowman; Vijay S. Pande (2010). "Protein folded states are kinetic hubs". Proceedings of the National Academy of Sciences. 107 (24): 10890–5. Bibcode:2010PNAS..10710890B. doi:10.1073/pnas.1003962107. PMC 2890711. PMID 20534497.
^ Jump up to: a b Christopher D. Snow; Houbi Nguyen; Vijay S. Pande; Martin Gruebele (2002). "Absolute comparison of simulated and experimental protein-folding dynamics" (PDF). Nature. 420 (6911): 102–106. Bibcode:2002Natur.420..102S. doi:10.1038/nature01160. PMID 12422224. Archived from the original (PDF) on March 24, 2012.
^ Fabrizio Marinelli, Fabio Pietrucci, Alessandro Laio, Stefano Piana (2009). Pande, Vijay S. (ed.). "A Kinetic Model of Trp-Cage Folding from Multiple Biased Molecular Dynamics Simulations". PLOS Computational Biology. 5 (8): e1000452. Bibcode:2009PLSCB...5E0452M. doi:10.1371/journal.pcbi.1000452. PMC 2711228. PMID 19662155.
^ "So Much More to Know". Science. 309 (5731): 78–102. 2005. doi:10.1126/science.309.5731.78b. PMID 15994524.
^ Jump up to: a b c Heath Ecroyd; John A. Carver (2008). "Unraveling the mysteries of protein folding and misfolding". IUBMB Life (review). 60 (12): 769–774. doi:10.1002/iub.117. PMID 18767168.
^ Jump up to: a b Yiwen Chen; Feng Ding; Huifen Nie; Adrian W. Serohijos; Shantanu Sharma; Kyle C. Wilcox; Shuangye Yin; Nikolay V. Dokholyan (2008). "Protein folding: Then and now". Archives of Biochemistry and Biophysics. 469 (1): 4–19. doi:10.1016/j.abb.2007.05.014. PMC 2173875. PMID 17585870.
^ Jump up to: a b Leila M Luheshi; Damian Crowther; Christopher Dobson (2008). "Protein misfolding and disease: from the test tube to the organism". Current Opinion in Chemical Biology. 12 (1): 25–31. doi:10.1016/j.cbpa.2008.02.011. PMID 18295611.
^ C. D. Snow; E. J. Sorin; Y. M. Rhee; V. S. Pande. (2005). "How well can simulation predict protein folding kinetics and thermodynamics?". Annual Review of Biophysics (review). 34: 43–69. doi:10.1146/annurev.biophys.34.040204.144447. PMID 15869383.
^ A. Verma; S.M. Gopal; A. Schug; J.S. Oh; K.V. Klenin; K.H. Lee; W. Wenzel (2008). Massively Parallel All Atom Protein Folding in a Single Day. Advances in Parallel Computing. 15. pp. 527–534. ISBN 978-1-58603-796-3. ISSN 0927-5452.
^ Vijay S. Pande; Ian Baker; Jarrod Chapman; Sidney P. Elmer; Siraj Khaliq; Stefan M. Larson; Young Min Rhee; Michael R. Shirts; Christopher D. Snow; Eric J. Sorin; Bojan Zagrovic (2002). "Atomistic protein folding simulations on the submillisecond timescale using worldwide distributed computing". Biopolymers. 68 (1): 91–109. doi:10.1002/bip.10219. PMID 12579582.
^ Jump up to: a b c G. Bowman; V. Volez; V. S. Pande (2011). "Taming the complexity of protein folding". Current Opinion in Structural Biology. 21 (1): 4–11. doi:10.1016/j.sbi.2010.10.006. PMC 3042729. PMID 21081274.
^ Chodera, John D.; Swope, William C.; Pitera, Jed W.; Dill, Ken A. (January 1, 2006). "Long‐Time Protein Folding Dynamics from Short‐Time Molecular Dynamics Simulations". Multiscale Modeling & Simulation. 5 (4): 1214–1226. doi:10.1137/06065146X.
^ Robert B Best (2012). "Atomistic molecular simulations of protein folding". Current Opinion in Structural Biology (review). 22 (1): 52–61. doi:10.1016/j.sbi.2011.12.001. PMID 22257762.
^ Jump up to: a b c TJ Lane; Gregory Bowman; Robert McGibbon; Christian Schwantes; Vijay Pande; Bruce Borden (September 10, 2012). "Folding@home Simulation FAQ". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved July 8, 2013.
^ Gregory R. Bowman; Daniel L. Ensign; Vijay S. Pande (2010). "Enhanced Modeling via Network Theory: Adaptive Sampling of Markov State Models". Journal of Chemical Theory and Computation. 6 (3): 787–794. doi:10.1021/ct900620b. PMC 3637129. PMID 23626502.
^ Vijay Pande (June 8, 2012). "FAHcon 2012: Thinking about how far FAH has come". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved June 12, 2012.
^ Kyle A. Beauchamp; Daniel L. Ensign; Rhiju Das; Vijay S. Pande (2011). "Quantitative comparison of villin headpiece subdomain simulations and triplet–triplet energy transfer experiments". Proceedings of the National Academy of Sciences. 108 (31): 12734–9. Bibcode:2011PNAS..10812734B. doi:10.1073/pnas.1010880108. PMC 3150881. PMID 21768345.
^ Timothy H. Click; Debabani Ganguly; Jianhan Chen (2010). "Intrinsically Disordered Proteins in a Physics-Based World". International Journal of Molecular Sciences. 11 (12): 919–27. doi:10.3390/ijms11125292. PMC 3100817. PMID 21614208.
^ "Greg Bowman awarded the 2010 Kuhn Paradigm Shift Award". simtk.org. SimTK: MSMBuilder. March 29, 2010. Archived from the original on September 21, 2012. Retrieved September 20,2012.
^ "MSMBuilder Source Code Repository". MSMBuilder. simtk.org. 2012. Archived from the original on October 12, 2012. Retrieved October 12, 2012.
^ "Biophysical Society Names Five 2012 Award Recipients". Biophysics.org. Biophysical Society. August 17, 2011. Archived from the original on September 21, 2012. Retrieved September 20,2012.
^ "Folding@home – Awards". Folding@home. foldingathome.org. August 2011. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ Vittorio Bellotti; Monica Stoppini (2009). "Protein Misfolding Diseases" (PDF). The Open Biology Journal. 2 (2): 228–234. doi:10.2174/1874196700902020228. Archived from the original on February 22, 2014.
^ Jump up to: a b c d e f g h i Pande lab (May 30, 2012). "Folding@home Diseases Studied FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ Jump up to: a b Collier, Leslie; Balows, Albert; Sussman, Max (1998). Mahy, Brian; Collier, Leslie (eds.). Topley and Wilson's Microbiology and Microbial Infections. 1, Virology (ninth ed.). London: Arnold. pp. 75–91. ISBN 978-0-340-66316-5.
^ Fred E. Cohen; Jeffery W. Kelly (2003). "Therapeutic approaches to protein misfolding diseases". Nature (review). 426 (6968): 905–9. Bibcode:2003Natur.426..905C. doi:10.1038/nature02265. PMID 14685252.
^ Jump up to: a b Chun Song; Shen Lim; Joo Tong (2009). "Recent advances in computer-aided drug design". Briefings in Bioinformatics (review). 10 (5): 579–91. doi:10.1093/bib/bbp023. PMID 19433475.
^ Jump up to: a b c d e f Pande lab (2012). "Folding@Home Press FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ Christian "schwancr" Schwantes (Pande lab member) (August 15, 2011). "Projects 7808 and 7809 to full fah". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved October 16, 2011.
^ Del Lucent; V. Vishal; Vijay S. Pande (2007). "Protein folding under confinement: A role for solvent". Proceedings of the National Academy of Sciences of the United States of America. 104(25): 10430–10434. Bibcode:2007PNAS..10410430L. doi:10.1073/pnas.0608256104. PMC 1965530. PMID 17563390.
^ Vincent A. Voelz; Vijay R. Singh; William J. Wedemeyer; Lisa J. Lapidus; Vijay S. Pande (2010). "Unfolded-State Dynamics and Structure of Protein L Characterized by Simulation and Experiment". Journal of the American Chemical Society. 132 (13): 4702–4709. doi:10.1021/ja908369h. PMC 2853762. PMID 20218718.
^ Jump up to: a b Vijay Pande (April 23, 2008). "Folding@home and Simbios". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved November 9, 2011.
^ Vijay Pande (October 25, 2011). "Re: Suggested Changes to F@h Website". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved October 25, 2011.
^ Caroline Hadley (2004). "Biologists think bigger". EMBO Reports. 5 (3): 236–238. doi:10.1038/sj.embor.7400108. PMC 1299019. PMID 14993921.
^ S. Pronk; P. Larsson; I. Pouya; G.R. Bowman; I.S. Haque; K. Beauchamp; B. Hess; V.S. Pande; P.M. Kasson; E. Lindahl (2011). "Copernicus: A new paradigm for parallel adaptive molecular dynamics". 2011 International Conference for High Performance Computing, Networking, Storage and Analysis: 1–10, 12–18.
^ Sander Pronk; Iman Pouya; Per Larsson; Peter Kasson; Erik Lindahl (November 17, 2011). "Copernicus Download". copernicus-computing.org. Copernicus. Archived from the original on October 12, 2012. Retrieved October 2, 2012.
^ Pande lab (July 27, 2012). "Papers & Results from Folding@home". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved 1 February 2019.
^ G Brent Irvine; Omar M El-Agnaf; Ganesh M Shankar; Dominic M Walsh (2008). "Protein Aggregation in the Brain: The Molecular Basis for Alzheimer's and Parkinson's Diseases". Molecular Medicine (review). 14 (7–8): 451–464. doi:10.2119/2007-00100.Irvine. PMC 2274891. PMID 18368143.
^ Claudio Soto; Lisbell D. Estrada (2008). "Protein Misfolding and Neurodegeneration". Archives of Neurology (review). 65 (2): 184–189. doi:10.1001/archneurol.2007.56. PMID 18268186.
^ Robin Roychaudhuri; Mingfeng Yang; Minako M. Hoshi; David B. Teplow (2008). "Amyloid β-Protein Assembly and Alzheimer Disease". Journal of Biological Chemistry. 284 (8): 4749–53. doi:10.1074/jbc.R800036200. PMC 3837440. PMID 18845536.
^ Jump up to: a b Nicholas W. Kelley; V. Vishal; Grant A. Krafft; Vijay S. Pande. (2008). "Simulating oligomerization at experimental concentrations and long timescales: A Markov state model approach". Journal of Chemical Physics. 129 (21): 214707. Bibcode:2008JChPh.129u4707K. doi:10.1063/1.3010881. PMC 2674793. PMID 19063575.
^ Jump up to: a b P. Novick, J. Rajadas, C.W. Liu, N. W. Kelley, M. Inayathullah, and V. S. Pande (2011). Buehler, Markus J. (ed.). "Rationally Designed Turn Promoting Mutation in the Amyloid-β Peptide Sequence Stabilizes Oligomers in Solution". PLOS ONE. 6 (7): e21776. Bibcode:2011PLoSO...621776R. doi:10.1371/journal.pone.0021776. PMC 3142112. PMID 21799748.
^ Aabgeena Naeem; Naveed Ahmad Fazili (2011). "Defective Protein Folding and Aggregation as the Basis of Neurodegenerative Diseases: The Darker Aspect of Proteins". Cell Biochemistry and Biophysics (review). 61 (2): 237–50. doi:10.1007/s12013-011-9200-x. PMID 21573992.
^ Jump up to: a b c d Gregory R Bowman; Xuhui Huang; Vijay S Pande (2010). "Network models for molecular kinetics and their initial applications to human health". Cell Research (review). 20 (6): 622–630. doi:10.1038/cr.2010.57. PMC 4441225. PMID 20421891.
^ Vijay Pande (December 18, 2008). "New FAH results on possible new Alzheimer's drug presented". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 23, 2011.
^ Paul A. Novick; Dahabada H. Lopes; Kim M. Branson; Alexandra Esteras-Chopo; Isabella A. Graef; Gal Bitan; Vijay S. Pande (2012). "Design of β-Amyloid Aggregation Inhibitors from a Predicted Structural Motif". Journal of Medicinal Chemistry. 55 (7): 3002–10. doi:10.1021/jm201332p. PMC 3766731. PMID 22420626.
^ yslin (Pande lab member) (July 22, 2011). "New project p6871 [Classic]". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved March 17, 2012.(registration required)
^ Pande lab. "Project 6871 Description". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved September 27, 2011.
^ Walker FO (2007). "Huntington's disease". Lancet. 369 (9557): 218–28 [220]. doi:10.1016/S0140-6736(07)60111-1. PMID 17240289.
^ Nicholas W. Kelley; Xuhui Huang; Stephen Tam; Christoph Spiess; Judith Frydman; Vijay S. Pande (2009). "The predicted structure of the headpiece of the Huntingtin protein and its implications on Huntingtin aggregation". Journal of Molecular Biology. 388 (5): 919–27. doi:10.1016/j.jmb.2009.01.032. PMC 2677131. PMID 19361448.
^ Susan W Liebman; Stephen C Meredith (2010). "Protein folding: Sticky N17 speeds huntingtin pile-up". Nature Chemical Biology. 6 (1): 7–8. doi:10.1038/nchembio.279. PMID 20016493.
^ Diwakar Shukla (Pande lab member) (February 10, 2012). "Project 8021 released to beta". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved March 17, 2012.(registration required)
^ M Hollstein; D Sidransky; B Vogelstein; CC Harris (1991). "p53 mutations in human cancers". Science. 253 (5015): 49–53. Bibcode:1991Sci...253...49H. doi:10.1126/science.1905840. PMID 1905840.
^ L. T. Chong; C. D. Snow; Y. M. Rhee; V. S. Pande. (2004). "Dimerization of the p53 Oligomerization Domain: Identification of a Folding Nucleus by Molecular Dynamics Simulations". Journal of Molecular Biology. 345 (4): 869–878. CiteSeerX 10.1.1.132.1174. doi:10.1016/j.jmb.2004.10.083. PMID 15588832.
^ mah3, Vijay Pande (September 24, 2004). "F@H project publishes results of cancer related research". MaximumPC.com. Future US, Inc. Archived from the original on September 21, 2012. Retrieved September 20, 2012. To our knowledge, this is the first peer-reviewed results from a distributed computing project related to cancer.
^ Lillian T. Chong; William C. Swope; Jed W. Pitera; Vijay S. Pande (2005). "Kinetic Computational Alanine Scanning: Application to p53 Oligomerization". Journal of Molecular Biology. 357 (3): 1039–1049. doi:10.1016/j.jmb.2005.12.083. PMID 16457841.
^ Almeida MB, do Nascimento JL, Herculano AM, Crespo-López ME (2011). "Molecular chaperones: toward new therapeutic tools". Journal of Molecular Biology (review). 65 (4): 239–43. doi:10.1016/j.biopha.2011.04.025. PMID 21737228.
^ Vijay Pande (September 28, 2007). "Nanomedicine center". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 23, 2011.
^ Vijay Pande (December 22, 2009). "Release of new Protomol (Core B4) WUs". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 23, 2011.
^ Pande lab. "Project 180 Description". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved September 27, 2011.
^ TJ Lane (Pande lab member) (June 8, 2011). "Project 7600 in Beta". Folding@home. phpBBGroup. Archived from the original on September 21, 2012. Retrieved September 27, 2011.(registration required)
^ TJ Lane (Pande lab member) (June 8, 2011). "Project 7600 Description". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved March 31, 2012.
^ "Scientists boost potency, reduce side effects of IL-2 protein used to treat cancer". MedicalXpress.com. Medical Xpress. March 18, 2012. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Aron M. Levin; Darren L. Bates; Aaron M. Ring; Carsten Krieg; Jack T. Lin; Leon Su; Ignacio Moraga; Miro E. Raeber; Gregory R. Bowman; Paul Novick; Vijay S. Pande; C. Garrison Fathman; Onur Boyman; K. Christopher Garcia (2012). "Exploiting a natural conformational switch to engineer an interleukin-2 'superkine'". Nature. 484 (7395): 529–33. Bibcode:2012Natur.484..529L. doi:10.1038/nature10975. PMC 3338870. PMID 22446627.
^ Rauch F, Glorieux FH (2004). "Osteogenesis imperfecta". Lancet. 363 (9418): 1377–85. doi:10.1016/S0140-6736(04)16051-0. PMID 15110498.
^ Fratzl, Peter (2008). Collagen: structure and mechanics. ISBN 978-0-387-73905-2. Retrieved March 17, 2012.
^ Gautieri A, Uzel S, Vesentini S, Redaelli A, Buehler MJ (2009). "Molecular and mesoscale disease mechanisms of Osteogenesis Imperfecta". Biophysical Journal. 97 (3): 857–865. Bibcode:2009BpJ....97..857G. doi:10.1016/j.bpj.2009.04.059. PMC 2718154. PMID 19651044.
^ Sanghyun Park; Randall J. Radmer; Teri E. Klein; Vijay S. Pande (2005). "A New Set of Molecular Mechanics Parameters for Hydroxyproline and Its Use in Molecular Dynamics Simulations of Collagen-Like Peptides". Journal of Computational Chemistry. 26 (15): 1612–1616. CiteSeerX 10.1.1.142.6781. doi:10.1002/jcc.20301. PMID 16170799.
^ Gregory Bowman (Pande lab Member). "Project 10125". Folding@home. phpBB Group. Retrieved December 2, 2011.(registration required)
^ Hana Robson Marsden; Itsuro Tomatsu; Alexander Kros (2011). "Model systems for membrane fusion". Chemical Society Reviews (review). 40 (3): 1572–1585. doi:10.1039/c0cs00115e. PMID 21152599.
^ Peter Kasson (2012). "Peter M. Kasson". Kasson lab. University of Virginia. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Peter M. Kasson; Afra Zomorodian; Sanghyun Park; Nina Singhal; Leonidas J. Guibas; Vijay S. Pande (2007). "Persistent voids: a new structural metric for membrane fusion". Bioinformatics. 23(14): 1753–1759. doi:10.1093/bioinformatics/btm250. PMID 17488753.
^ Peter M. Kasson; Daniel L. Ensign; Vijay S. Pande (2009). "Combining Molecular Dynamics with Bayesian Analysis To Predict and Evaluate Ligand-Binding Mutations in Influenza Hemagglutinin". Journal of the American Chemical Society. 131 (32): 11338–11340. doi:10.1021/ja904557w. PMC 2737089. PMID 19637916.
^ Peter M. Kasson; Vijay S. Pande (2009). "Combining mutual information with structural analysis to screen for functionally important residues in influenza hemagglutinin". Pacific Symposium on Biocomputing: 492–503. doi:10.1142/9789812836939_0047. ISBN 978-981-283-692-2. PMC 2811693. PMID 19209725.
^ Vijay Pande (February 24, 2012). "Protein folding and viral infection". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved March 4, 2012.
^ Broekhuijsen, Niels (March 3, 2020). "Help Cure Coronavirus with Your PC's Leftover Processing Power". Tom's Hardware. Retrieved March 12, 2020.
^ Bowman, Greg (February 27, 2020). "Folding@home takes up the fight against COVID-19 / 2019-nCoV". Folding@home. Retrieved March 12, 2020.
^ "Folding@home Turns Its Massive Crowdsourced Computer Network Against COVID-19". March 16, 2020.
^ Vijay Pande (February 27, 2012). "New methods for computational drug design". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved April 1, 2012.
^ Guha Jayachandran; M. R. Shirts; S. Park; V. S. Pande (2006). "Parallelized-Over-Parts Computation of Absolute Binding Free Energy with Docking and Molecular Dynamics". Journal of Chemical Physics. 125 (8): 084901. Bibcode:2006JChPh.125h4901J. doi:10.1063/1.2221680. PMID 16965051.
^ Pande lab. "Project 10721 Description". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved September 27, 2011.
^ Jump up to: a b Gregory Bowman (July 23, 2012). "Searching for new drug targets". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 27, 2011.
^ Gregory R. Bowman; Phillip L. Geissler (July 2012). "Equilibrium fluctuations of a single folded protein reveal a multitude of potential cryptic allosteric sites". PNAS. 109 (29): 11681–6. Bibcode:2012PNAS..10911681B. doi:10.1073/pnas.1209309109. PMC 3406870. PMID 22753506.
^ Paula M. Petrone; Christopher D. Snow; Del Lucent; Vijay S. Pande (2008). "Side-chain recognition and gating in the ribosome exit tunnel". Proceedings of the National Academy of Sciences. 105 (43): 16549–54. Bibcode:2008PNAS..10516549P. doi:10.1073/pnas.0801795105. PMC 2575457. PMID 18946046.
^ Pande lab. "Project 5765 Description". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved December 2, 2011.
^ National Institute of Neurological Disorders and Stroke (August 21, 2018). "Creutzfeldt-Jakob Disease Fact Sheet". NIH. Retrieved March 2, 2019.
^ Jump up to: a b c d e f Kupfer, L; Hinrichs, W; Groschup, M.H. (2009). "Prion Protein Misfolding". Current Molecular Medicine. Bentham Science Publishers. 9 (7): 826–835. doi:10.2174/156652409789105543. PMC 3330701. PMID 19860662.
^ Jump up to: a b c Riek, Poland; Hornemann, Simone; Wider, Gerhard; Billeter, Martin; Glockshuber, Rudi; Wüthrich, Kurt (1996). "NMR structure of the mouse prion protein domain in PrP(121-231)". Nature. Nature Research. 382 (6587): 180–182. Bibcode:1996Natur.382..180R. doi:10.1038/382180a0. PMID 8700211.
^ Jump up to: a b Ziegler, J; Sticht, H; Marx, UC; Müller, W; Rösch, P; Schwarzinger, S (2003). "CD and NMR studies of prion protein (PrP) helix1. Novel implications for its role in the PrPC-->PrPSc conversion process" (PDF). J Biol Chem. American Society for Biochemistry and Molecular Biology. 278 (50): 50175–81. doi:10.1074/jbc.M305234200. PMID 12952977.
^ Govaerts, Cedric; Wile, Holger; Brusiner, Stanley B.; Cohen, Fred (2004). "Evidence for assembly of prions with left-handed β-helices into trimers". Proc Natl Acad Sci USA. National Academy of Sciences. 101 (22): 8342–47. Bibcode:2004PNAS..101.8342G. doi:10.1073/pnas.0402254101. PMC 420396. PMID 15155909.
^ Silveira, Jay; Raymond, Gregory; Hughson, Andrew; Race, Richard; Sim, Valerie; Caughey, Byron; Hayes, Stanley (2005). "The most infectious prion protein particles". Nature. Nature Research. 437 (7056): 257–261. Bibcode:2005Natur.437..257S. doi:10.1038/nature03989. PMC 1513539. PMID 16148934.
^ Jump up to: a b Moore, Roger A.; Taubner, Lara M.; Priola, Suzette (2009). "Prion Protein Misfolding and Disease". Curr Opin Struct Biol. Elsevier. 19 (1): 14–22. doi:10.1016/j.sbi.2008.12.007. PMC 2674794. PMID 19157856.
^ Moore, Roger A.; Herzog, Christian; Errett, John; Kocisko, David A.; Arnold, Kevin M.; Hayes, Stanley F.; Priola, Suzette A. (2006). "Octapeptide repeat insertions increase the rate of protease-resistant prion protein formation". Protein Science. Wiley-Blackwell. 15 (3): 609–619. doi:10.1110/ps.051822606. PMC 2249780. PMID 16452616.
^ Gasset, M; Baldwin, M; Lloyd, D; Gabriel, J; Holtzman, D; Cohen, F; Fletterick, R; Brusiner, S (1992). "Predicted alpha-helical regions of the prion protein when synthesized as peptides form amyloid". Proc Natl Acad Sci USA. National Academy of Sciences. 89 (22): 10940–44. Bibcode:1992PNAS...8910940G. doi:10.1073/pnas.89.22.10940. PMC 50458. PMID 1438300.
^ Ziegler, Jan; Viehrig, Christine; Geimer, Stefan; Rosch, Paul; Schwarzinger, Stephan (2006). "Putative aggregation initiation sites in prion protein". FEBS Letters. FEBS Press. 580 (8): 2033–40. doi:10.1016/j.febslet.2006.03.002. PMID 16545382.
^ Brusiner, Stanley (1998). "Prions". Proc Natl Acad Sci USA. National Academy of Sciences. 95(23): 13363–83. Bibcode:1998PNAS...9513363P. doi:10.1073/pnas.95.23.13363. PMC 33918. PMID 9811807.
^ Vorberg, I; Groschup, MH; Pfaff, E; Priola, SA (2003). "Multiple amino acid residues within the rabbit prion protein inhibit formation of its abnormal isoform". J. Virol. American Society for Microbiology. 77 (3): 2003–9. doi:10.1128/JVI.77.3.2003-2009.2003. PMC 140934. PMID 12525634.
^ Barrow, CJ; Yasuda, A; Kenny, PT; Zagorski, MG (1992). "Solution conformations and aggregational properties of synthetic amyloid beta-peptides of Alzheimer's disease. Analysis of circular dichroism spectra". J Biol Chem. American Society for Biochemistry and Molecular Biology. 225 (4): 1075–93. doi:10.1016/0022-2836(92)90106-t. PMID 1613791.
^ Kuznetsov, Igor; Rackovsky, Shalom (2004). "Comparative computational analysis of prion proteins reveals two fragments with unusual structural properties and a pattern of increase in hydrophobicity associated with disease-promoting mutations". Protein Science. Wiley-Blackwell. 13 (12): 3230–44. doi:10.1110/ps.04833404. PMC 2287303. PMID 15557265.
^ Baskakov, IV; Legname, G; Baldwin, MA; Prusiner, SB; Cohen, FE (2002). "Pathway complexity of prion protein assembly into amyloid". J Biol Chem. American Society for Biochemistry and Molecular Biology. 277 (24): 21140–8. doi:10.1074/jbc.M111402200. PMID 11912192.
^ Raddick, M. Jordan; Bracey, Georgia; Gay, Pamela L.; Lintott, Chris J.; Murray, Phil; Schawinski, Kevin; Szalay, Alexander S.; Vandenberg, Jan (December 2010). "Galaxy Zoo: Exploring the Motivations of Citizen Science Volunteers". Astronomy Education Review. 9 (1): 010103. arXiv:0909.2925. Bibcode:2010AEdRv...9a0103R. doi:10.3847/AER2009036.
^ Vickie, Curtis (April 20, 2018). Online citizen science and the widening of academia : distributed engagement with research and knowledge production. Cham, Switzerland. ISBN 9783319776644. OCLC 1034547418.
^ Nov, Oded; Arazy, Ofer; Anderson, David (2011). "Dusting for science: motivation and participation of digital citizen science volunteers". Proceedings of the 2011 IConference on - IConference '11. IConference '11. Seattle, Washington: ACM Press: 68–74. doi:10.1145/1940761.1940771. ISBN 9781450301213.
^ Curtis, Vickie (December 2015). "Motivation to Participate in an Online Citizen Science Game: A Study of Foldit" (PDF). Science Communication. 37 (6): 723–746. doi:10.1177/1075547015609322. ISSN 1075-5470.
^ Jump up to: a b c Curtis, Vickie (April 27, 2018). "Patterns of Participation and Motivation in Folding@home: The Contribution of Hardware Enthusiasts and Overclockers". Citizen Science: Theory and Practice. 3 (1): 5. doi:10.5334/cstp.109. ISSN 2057-4991.
^ Colwell, B. (March 2004). "The Zen of overclocking". Computer. 37 (3): 9–12. doi:10.1109/MC.2004.1273994. ISSN 0018-9162.
^ Kloetzer, Laure; Da Costa, Julien; Schneider, Daniel K. (December 31, 2016). "Not so passive: engagement and learning in Volunteer Computing projects". Human Computation. 3 (1). doi:10.15346/hc.v3i1.4. ISSN 2330-8001.
^ Darch Peter; Carusi Annamaria (September 13, 2010). "Retaining volunteers in volunteer computing projects". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 368 (1926): 4177–4192. Bibcode:2010RSPTA.368.4177D. doi:10.1098/rsta.2010.0163. PMID 20679130.
^ "2013 Member Study: Findings and Next Steps". World Community Grid.
^ Krebs, Viola (January 31, 2010). "Motivations of cybervolunteers in an applied distributed computing environment: MalariaControl.net as an example". First Monday. 15 (2). doi:10.5210/fm.v15i2.2783.
^ Curtis, Vickie (2015). Online citizen science projects: an exploration of motivation, contribution and participation, PhD Thesis (PDF). United Kingdom: The Open University.
^ Mims 2010
^ Pande 2008: "Wall clock time is in the end the only thing that matters and that's why we benchmark on wall clock (and why in our papers, we emphasize wall clock)."
^ Vijay Pande (September 16, 2007). "Crossing the petaFLOPS barrier". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved August 28, 2011.
^ Jump up to: a b Michael Gross (2012). "Folding research recruits unconventional help". Current Biology. 22 (2): R35–R38. doi:10.1016/j.cub.2012.01.008. PMID 22389910.
^ "TOP500 List — June 2007". top500.org. Top500. June 2007. Retrieved September 20, 2012.
^ "Folding@Home reach 2 Petaflops". n4g.com. HAVAmedia. May 8, 2008. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ "NVIDIA Achieves Monumental Folding@Home Milestone With Cuda". nvidia.com. NVIDIA Corporation. August 26, 2008. Retrieved September 20, 2012.
^ "3 PetaFLOP barrier". longecity.org. Longecity. August 19, 2008. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ "Increase in 'active' PS3 folders pushes Folding@home past 4 Petaflops!". team52735.blogspot.com. Blogspot. September 29, 2008. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Vijay Pande (February 18, 2009). "Folding@home Passes the 5 petaFLOP Mark". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved August 31, 2011.
^ "Crossing the 5 petaFLOPS barrier". longecity.org. Longecity. February 18, 2009. Archivedfrom the original on September 21, 2012. Retrieved September 20, 2012.
^ Dragan Zakic (May 2009). "Community Grid Computing — Studies in Parallel and Distributed Systems" (PDF). Massey University College of Sciences. Massey University. Archived from the original (PDF) on September 21, 2012. Retrieved September 20, 2012.
^ William Ito. "A review of recent advances in ab initio protein folding by the Folding@home project" (PDF). foldingathome.org. Archived from the original (PDF) on September 22, 2012. Retrieved September 22, 2012.
^ "TOP500 List — November 2008". top500.org. Top500. November 2008. Retrieved September 20, 2012.
^ Jesse Victors (November 10, 2011). "Six Native PetaFLOPS". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved November 11, 2011.
^ Risto Kantonen (September 23, 2013). "Folding@home Stats - Google Docs". Folding@home. Retrieved September 23, 2013.
^ "100 Petaflops nearly reached". foldingathome.org. May 11, 2016. Retrieved August 9, 2016.
^ Bowman, Greg (March 20, 2020). "Amazing! @foldingathome now has over 470 petaFLOPS of compute power. To put that in perspective, that's more than 2x the peak performance of the Summit super computer!". @drGregBowman. Retrieved March 20, 2020.
^ "Folding@home stats report". March 20, 2020. Archived from the original on March 20, 2020. Retrieved March 20, 2020.
^ Shilov, Anton (March 25, 2020). "Folding@Home Reaches Exascale: 1,500,000,000,000,000,000 Operations Per Second for COVID-19". Anandtech. Retrieved March 26, 2020.
^ Jump up to: a b c d e Pande lab (August 20, 2012). "Folding@home Points FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8,2013.
^ Pande lab (July 23, 2012). "Folding@home Passkey FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ Jump up to: a b Peter Kasson (Pande lab member) (January 24, 2010). "upcoming release of SMP2 cores". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved September 30, 2011.
^ Jump up to: a b c d e Pande lab (August 18, 2011). "Folding@home Main FAQ" (FAQ). Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved July 8, 2013.
^ "Official Extreme Overclocking Folding@home Team Forum". forums.extremeoverclocking.com. Extreme Overclocking. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Jump up to: a b c d e f g Adam Beberg; Daniel Ensign; Guha Jayachandran; Siraj Khaliq; Vijay Pande (2009). "Folding@home: Lessons from eight years of volunteer distributed computing" (PDF). 2009 IEEE International Symposium on Parallel & Distributed Processing. Proceedings. pp. 1–8. doi:10.1109/IPDPS.2009.5160922. ISBN 978-1-4244-3751-1. ISSN 1530-2075.
^ Norman Chan (April 6, 2009). "Help Maximum PC's Folding Team Win the Next Chimp Challenge!". Maximumpc.com. Future US, Inc. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Jump up to: a b c Pande lab (June 11, 2012). "Folding@home SMP FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8,2013.
^ Vijay Pande (April 5, 2011). "More transparency in testing". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved October 14, 2011.
^ Bruce Borden (August 7, 2011). "Re: Gromacs Cannot Continue Further". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved August 7, 2011.
^ PantherX (October 1, 2011). "Re: Project 6803: (Run 4, Clone 66, Gen 255)". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved October 9, 2011.
^ PantherX (October 31, 2010). "Troubleshooting Bad WUs". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved August 7, 2011.
^ Carsten Kutzner; David Van Der Spoel; Martin Fechner; Erik Lindahl; Udo W. Schmitt; Bert L. De Groot; Helmut Grubmüller (2007). "Speeding up parallel GROMACS on high-latency networks". Journal of Computational Chemistry. 28 (12): 2075–2084. doi:10.1002/jcc.20703. hdl:11858/00-001M-0000-0012-E29A-0. PMID 17405124.
^ Berk Hess; Carsten Kutzner; David van der Spoel; Erik Lindahl (2008). "GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation". Journal of Chemical Theory and Computation. 4 (3): 435–447. doi:10.1021/ct700301q. hdl:11858/00-001M-0000-0012-DDBF-0. PMID 26620784.
^ Pande lab (August 19, 2012). "Folding@home Gromacs FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8,2013.
^ Pande lab (August 7, 2012). "Folding@home Frequently Asked Questions (FAQ) Index". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved July 8, 2013.
^ Vijay Pande (September 25, 2009). "Update on new FAH cores and clients". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved February 24, 2012.
^ Jump up to: a b M. S. Friedrichs; P. Eastman; V. Vaidyanathan; M. Houston; S. LeGrand; A. L. Beberg; D. L. Ensign; C. M. Bruns; V. S. Pande (2009). "Accelerating Molecular Dynamic Simulation on Graphics Processing Units". Journal of Computational Chemistry. 30 (6): 864–72. doi:10.1002/jcc.21209. PMC 2724265. PMID 19191337.
^ Pande lab (August 19, 2012). "Folding@home Petaflop Initiative". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8,2013.
^ Jump up to: a b Pande lab (February 10, 2011). "Windows Uniprocessor Client Installation Guide". Folding@home. foldingathome.org. Archived from the original (Guide) on September 21, 2012. Retrieved July 8, 2013.
^ PantherX (September 2, 2010). "Re: Can Folding@home damage any part of my PC?". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved February 25, 2012.
^ Jump up to: a b c Vijay Pande (June 17, 2009). "How does FAH code development and sysadmin get done?". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved October 14, 2011.
^ Jump up to: a b Pande lab (May 30, 2012). "Uninstalling Folding@home Guide". Folding@home. foldingathome.org. Archived from the original (Guide) on September 21, 2012. Retrieved July 8,2013.
^ Jump up to: a b Folding@home developers. "FAHControl source code repository". foldingathome.org. Archived from the original on December 12, 2012. Retrieved October 15, 2012.
^ Pande lab. "Folding@home Distributed Computing Client". Folding@home. foldingathome.org. Archived from the original on September 21, 2012. Retrieved July 8, 2013.
^ Vijay Pande (June 28, 2008). "Folding@home's End User License Agreement (EULA)". Folding@home. Archived from the original on September 21, 2012. Retrieved May 15, 2012.
^ unikuser (August 7, 2011). "FoldingAtHome". Ubuntu Documentation. help.ubuntu.com. Archived from the original on September 22, 2012. Retrieved September 22, 2012.
^ "The Case of the Modified Binaries". Leviathan Security.
^ "Fixing/Making Holes in ELF Binaries/Programs - Black Hat".
^ probably using tools such as ERESI
^ "x86 - How to disassemble, modify and then reassemble a Linux executable?". Stack Overflow.
^ "linux - How do I add functionality to an existing binary executable?". Reverse Engineering Stack Exchange.
^ "Certificate Bypass: Hiding and Executing Malware from a Digitally Signed Executable" (PDF). BlackHat.com. Deep Instinct. August 2016.
^ Phineus R. L. Markwick; J. Andrew McCammon (2011). "Studying functional dynamics in bio-molecules using accelerated molecular dynamics". Physical Chemistry Chemical Physics. 13 (45): 20053–65. Bibcode:2011PCCP...1320053M. doi:10.1039/C1CP22100K. PMID 22015376.
^ M. R. Shirts; V. S. Pande. (2000). "Screen Savers of the World, Unite!". Science. 290 (5498): 1903–1904. doi:10.1126/science.290.5498.1903. PMID 17742054.
^ Pande lab. "Folding@Home Executive summary" (PDF). Folding@home. foldingathome.org. Archived (PDF) from the original on September 21, 2012. Retrieved October 4, 2011.
^ Rattledagger, Vijay Pande (April 1, 2005). "Folding@home client for BOINC in beta "soon"". Boarddigger.com. Anandtech.com. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Jump up to: a b c Pande lab (May 30, 2012). "High Performance FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ John D. Owens; David Luebke; Naga Govindaraju; Mark Harris; Jens Krüger; Aaron Lefohn; Timothy J. Purcell (2007). "A Survey of General-Purpose Computation on Graphics Hardware". Computer Graphics Forum. 26 (1): 80–113. CiteSeerX 10.1.1.215.426. doi:10.1111/j.1467-8659.2007.01012.x.
^ P. Eastman; V. S. Pande (2010). "OpenMM: A Hardware-Independent Framework for Molecular Simulations". Computing in Science and Engineering. 12 (4): 34–39. Bibcode:2010CSE....12d..34E. doi:10.1109/MCSE.2010.27. ISSN 1521-9615. PMC 4486654. PMID 26146490.
^ I. Haque; V. S. Pande (2010). "Hard Data on Soft Errors: A Large-Scale Assessment of Real-World Error Rates in GPGPU". 2010 10th IEEE/ACM International Conference on Cluster, Cloud and Grid Computing. pp. 691–696. arXiv:0910.0505. doi:10.1109/CCGRID.2010.84. ISBN 978-1-4244-6987-1.
^ Jump up to: a b c Pande lab (March 18, 2011). "ATI FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ Vijay Pande (May 23, 2008). "GPU news (about GPU1, GPU2, & NVIDIA support)". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 8, 2011.
^ Travis Desell; Anthony Waters; Malik Magdon-Ismail; Boleslaw K. Szymanski; Carlos A. Varela; Matthew Newby; Heidi Newberg; Andreas Przystawik; David Anderson (2009). "Accelerating the MilkyWay@Home volunteer computing project with GPUs". 8th International Conference on Parallel Processing and Applied Mathematics (PPAM 2009) Part I. pp. 276–288. CiteSeerX 10.1.1.158.7614. ISBN 978-3-642-14389-2.
^ Vijay Pande (April 10, 2008). "GPU2 open beta". Folding@home. typepad.com. Archivedfrom the original on September 21, 2012. Retrieved September 7, 2011.
^ Vijay Pande (April 15, 2008). "Updates to the Download page/GPU2 goes live". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 7, 2011.
^ Vijay Pande (April 11, 2008). "GPU2 open beta going well". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 7, 2011.
^ Jump up to: a b Vijay Pande (April 24, 2010). "Prepping for the GPU3 rolling: new client and NVIDIA FAH GPU clients will (in the future) need CUDA 2.2 or later". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 8, 2011.
^ Vijay Pande (May 25, 2010). "Folding@home: Open beta release of the GPU3 client/core". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 7, 2011.
^ Joseph Coffland (CEO of Cauldron Development LLC & lead developer at Folding@home) (October 13, 2011). "Re: FAHClient V7.1.38 released (4th Open-Beta)". Folding@home. phpBBGroup. Archived from the original on September 21, 2012. Retrieved October 15, 2011.
^ "NVIDIA GPU3 Linux/Wine Headless Install Guide". Folding@home. phpBB Group. November 8, 2008. Archived from the original on September 21, 2012. Retrieved September 5, 2011.
^ Pande lab. "Client Statistics by OS". Folding@home. foldingathome.org. Archived from the original on November 28, 2012. Retrieved July 8, 2013.
^ Bruce Borden (June 25, 2013). "GPU FahCore_17 is now available on Windows & native Linux". Folding@home. phpBB Group. Retrieved September 30, 2014.
^ "Futures in Biotech 27: Folding@home at 1.3 Petaflops". Castroller.com. CastRoller. December 28, 2007. Archived from the original (Interview, webcast) on September 21, 2012. Retrieved September 20, 2012.
^ Jump up to: a b Edgar Luttmann; Daniel L. Ensign; Vishal Vaidyanathan; Mike Houston; Noam Rimon; Jeppe Øland; Guha Jayachandran; Mark Friedrichs; Vijay S. Pande (2008). "Accelerating Molecular Dynamic Simulation on the Cell processor and PlayStation 3". Journal of Computational Chemistry. 30 (2): 268–274. doi:10.1002/jcc.21054. PMID 18615421.
^ Jump up to: a b David E. Williams (October 20, 2006). "PlayStation's serious side: Fighting disease". CNN. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Jerry Liao (March 23, 2007). "The Home Cure: PlayStation 3 to Help Study Causes of Cancer". mb.com. Manila Bulletin Publishing Corporation. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Lou Kesten, Associated Press (March 26, 2007). "Week in video-game news: 'God of War II' storms the PS2; a PS3 research project". Post-Gazette.com. Pittsburgh Post-Gazette. Archivedfrom the original on September 21, 2012. Retrieved September 20, 2012.
^ Elaine Chow (September 18, 2008). "PS3 News Service, Life With Playstation, Now Up For Download". Gizmodo.com. Gizmodo. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Vijay Pande (September 18, 2008). "Life with Playstation – a new update to the FAH/PS3 client". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved February 24, 2012.
^ Pande lab (May 30, 2012). "PS3 FAQ". Folding@home. foldingathome.org. Archived from the original (FAQ) on September 21, 2012. Retrieved July 8, 2013.
^ Eric Lempel (October 21, 2012). "PS3 System Software Update (v4.30)". PlayStation blog. Sony. Archived from the original on October 22, 2012. Retrieved October 21, 2012.
^ "Termination of Life with PlayStation". Life with PlayStation. Sony. November 6, 2012. Archived from the original on November 9, 2012. Retrieved November 8, 2012.
^ Jump up to: a b Vijay Pande (June 15, 2008). "What does the SMP core do?". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 7, 2011.
^ Daniel L. Ensign; Peter M. Kasson; Vijay S. Pande (2007). "Heterogeneity Even at the Speed Limit of Folding: Large-scale Molecular Dynamics Study of a Fast-folding Variant of the Villin Headpiece". Journal of Molecular Biology. 374 (3): 806–816. doi:10.1016/j.jmb.2007.09.069. PMC 3689540. PMID 17950314.
^ Jump up to: a b Vijay Pande (March 8, 2008). "New Windows client/core development (SMP and classic clients)". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 30, 2011.
^ Jump up to: a b Peter Kasson (Pande lab member) (July 15, 2009). "new release: extra-large work units". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved October 9, 2011.
^ Vijay Pande (February 7, 2012). "Update on "bigadv-16", the new bigadv rollout". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved February 9, 2012.
^ Vijay Pande (July 2, 2011). "Change in the points system for bigadv work units". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved February 24, 2012.
^ Vijay Pande (January 15, 2014). "Revised plans for BigAdv (BA) experiment". Retrieved October 6, 2014.
^ Jump up to: a b Pande lab (March 23, 2012). "Windows (FAH V7) Installation Guide". Folding@home. foldingathome.org. Archived from the original (Guide) on September 21, 2012. Retrieved July 8,2013.
^ Jump up to: a b c Vijay Pande (March 29, 2011). "Client version 7 now in open beta". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved August 14, 2011.
^ Vijay Pande (March 22, 2012). "Web page revamp and v7 rollout". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved March 22, 2012.
^ Vijay Pande (March 31, 2011). "Core 16 for ATI released; also note on NVIDIA GPU support for older boards". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved September 7, 2011.
^ aschofield and jcoffland (October 3, 2011). "Ticket #736 (Link to GPL in FAHControl)". Folding@home. Trac. Archived from the original on October 12, 2012. Retrieved October 12,2012.
^ Pande, Vijay (February 24, 2014). "Adding a completely new way to fold, directly in the browser". foldingathome.org. Pande Lab, Stanford University. Retrieved February 13, 2015.
^ "NaCL Web Client Shutdown Notice". Folding@Home. Folding@Home. Retrieved August 29,2019.
^ Pande, Vijay (July 7, 2015). "First full version of our Folding@Home client for Android Mobile phones". Folding@Home. foldingathome.org. Retrieved May 31, 2016.
^ "Folding@Home". Google Play. 2016. Retrieved May 31, 2016.
^ "Android client overhaul". Folding@home. February 2, 2018. Retrieved July 22, 2019.
^ Lensink MF, Méndez R, Wodak SJ (December 2007). "Docking and scoring protein complexes: CAPRI 3rd Edition". Proteins. 69 (4): 704–18. doi:10.1002/prot.21804. PMID 17918726.
^ Gregory R. Bowman; Vijay S. Pande (2009). "Simulated tempering yields insight into the low-resolution Rosetta scoring function". Proteins: Structure, Function, and Bioinformatics. 74 (3): 777–88. doi:10.1002/prot.22210. PMID 18767152.
^ G. R. Bowman and V. S. Pande (2009). Hofmann, Andreas (ed.). "The Roles of Entropy and Kinetics in Structure Prediction". PLOS ONE. 4 (6): e5840. Bibcode:2009PLoSO...4.5840B. doi:10.1371/journal.pone.0005840. PMC 2688754. PMID 19513117.
^ Gen_X_Accord, Vijay Pande (June 11, 2006). "Folding@home vs. Rosetta@home". Rosetta@home forums. University of Washington. Archived from the original on September 21, 2012. Retrieved September 20, 2012.
^ Vijay Pande (October 13, 2011). "Comparison between FAH and Anton's approaches". Folding@home. typepad.com. Archived from the original on September 21, 2012. Retrieved February 25, 2012.
^ Jump up to: a b c Thomas J. Lane; Gregory R. Bowman; Kyle A Beauchamp; Vincent Alvin Voelz; Vijay S. Pande (2011). "Markov State Model Reveals Folding and Functional Dynamics in Ultra-Long MD Trajectories". Journal of the American Chemical Society. 133 (45): 18413–9. doi:10.1021/ja207470h. PMC 3227799. PMID 21988563.
^ David E. Shaw; et al. (2009). Millisecond-scale molecular dynamics simulations on Anton. Proceedings of the Conference on High Performance Computing Networking, Storage and Analysis. pp. 1–11. doi:10.1145/1654059.1654099. ISBN 978-1-60558-744-8.
^ David E. Shaw; et al. (2010). "Atomic-Level Characterization of the Structural Dynamics of Proteins". Science. 330 (6002): 341–346. Bibcode:2010Sci...330..341S. doi:10.1126/science.1187409. PMID 20947758.
^ David E. Shaw; Martin M. Deneroff; Ron O. Dror; Jeffrey S. Kuskin; Richard H. Larson; John K. Salmon; Cliff Young; Brannon Batson; Kevin J. Bowers; Jack C. Chao; Michael P. Eastwood; Joseph Gagliardo; J. P. Grossman; C. Richard Ho; Douglas J. Ierardi; et al. (2008). "Anton, A Special-Purpose Machine for Molecular Dynamics Simulation". Communications of the ACM. 51 (7): 91–97. doi:10.1145/1364782.1364802.
^ Ron O. Dror; Robert M. Dirks; J.P. Grossman; Huafeng Xu; David E. Shaw (2012). "Biomolecular Simulation: A Computational Microscope for Molecular Biology". Annual Review of Biophysics. 41: 429–52. doi:10.1146/annurev-biophys-042910-155245. PMID 22577825.
^ TJ Lane (Pande lab member) (June 6, 2011). "Project 7610 & 7611 in Beta". Folding@home. phpBB Group. Archived from the original on September 21, 2012. Retrieved February 25, 2012.(registration required)
^ Pande lab. "Project 7610 Description". Folding@home. Archived from the original on September 21, 2012. Retrieved February 26, 2012.
Sources[edit]
Folding@home (n.d.e), "About", Folding@home, retrieved April 26, 2020
Mims, Christopher (November 8, 2010), "Why China's New Supercomputer Is Only Technically the World's Fastest", Technology Review, MIT, archived from the original on September 21, 2012, retrieved September 20, 2012
Pande, Vijay S. (November 10, 2008), "Re: ATI and NVIDIA stats vs. PPD numbers", Folding Forum, the fifth post from below, archived from the original on September 21, 2012, retrieved April 26, 2020
External links[edit]
Wikimedia Commons has media related to Folding@home.

Official website
Listen to this article (info/dl)


MENU


0:00





This audio file was created from a revision of the article "Folding@home" dated 2014-10-07, and does not reflect subsequent edits to the article. (Audio help)
More spoken articles

show