1
|
Venter JC, Adams MD, Myers EW, Li PW, Mural RJ, Sutton GG, Smith HO, Yandell M, Evans CA, Holt RA, Gocayne JD, Amanatides P, Ballew RM, Huson DH, Wortman JR, Zhang Q, Kodira CD, Zheng XH, Chen L, Skupski M, Subramanian G, Thomas PD, Zhang J, Gabor Miklos GL, Nelson C, Broder S, Clark AG, Nadeau J, McKusick VA, Zinder N, Levine AJ, Roberts RJ, Simon M, Slayman C, Hunkapiller M, Bolanos R, Delcher A, Dew I, Fasulo D, Flanigan M, Florea L, Halpern A, Hannenhalli S, Kravitz S, Levy S, Mobarry C, Reinert K, Remington K, Abu-Threideh J, Beasley E, Biddick K, Bonazzi V, Brandon R, Cargill M, Chandramouliswaran I, Charlab R, Chaturvedi K, Deng Z, Di Francesco V, Dunn P, Eilbeck K, Evangelista C, Gabrielian AE, Gan W, Ge W, Gong F, Gu Z, Guan P, Heiman TJ, Higgins ME, Ji RR, Ke Z, Ketchum KA, Lai Z, Lei Y, Li Z, Li J, Liang Y, Lin X, Lu F, Merkulov GV, Milshina N, Moore HM, Naik AK, Narayan VA, Neelam B, Nusskern D, Rusch DB, Salzberg S, Shao W, Shue B, Sun J, Wang Z, Wang A, Wang X, Wang J, Wei M, Wides R, Xiao C, Yan C, Yao A, Ye J, Zhan M, Zhang W, Zhang H, Zhao Q, Zheng L, Zhong F, Zhong W, Zhu S, Zhao S, Gilbert D, Baumhueter S, Spier G, Carter C, Cravchik A, Woodage T, Ali F, An H, Awe A, Baldwin D, Baden H, Barnstead M, Barrow I, Beeson K, Busam D, Carver A, Center A, Cheng ML, Curry L, Danaher S, Davenport L, Desilets R, Dietz S, Dodson K, Doup L, Ferriera S, Garg N, Gluecksmann A, Hart B, Haynes J, Haynes C, Heiner C, Hladun S, Hostin D, Houck J, Howland T, Ibegwam C, Johnson J, Kalush F, Kline L, Koduru S, Love A, Mann F, May D, McCawley S, McIntosh T, McMullen I, Moy M, Moy L, Murphy B, Nelson K, Pfannkoch C, Pratts E, Puri V, Qureshi H, Reardon M, Rodriguez R, Rogers YH, Romblad D, Ruhfel B, Scott R, Sitter C, Smallwood M, Stewart E, Strong R, Suh E, Thomas R, Tint NN, Tse S, Vech C, Wang G, Wetter J, Williams S, Williams M, Windsor S, Winn-Deen E, Wolfe K, Zaveri J, Zaveri K, Abril JF, Guigó R, Campbell MJ, Sjolander KV, Karlak B, Kejariwal A, Mi H, Lazareva B, Hatton T, Narechania A, Diemer K, Muruganujan A, Guo N, Sato S, Bafna V, Istrail S, Lippert R, Schwartz R, Walenz B, Yooseph S, Allen D, Basu A, Baxendale J, Blick L, Caminha M, Carnes-Stine J, Caulk P, Chiang YH, Coyne M, Dahlke C, Deslattes Mays A, Dombroski M, Donnelly M, Ely D, Esparham S, Fosler C, Gire H, Glanowski S, Glasser K, Glodek A, Gorokhov M, Graham K, Gropman B, Harris M, Heil J, Henderson S, Hoover J, Jennings D, Jordan C, Jordan J, Kasha J, Kagan L, Kraft C, Levitsky A, Lewis M, Liu X, Lopez J, Ma D, Majoros W, McDaniel J, Murphy S, Newman M, Nguyen T, Nguyen N, Nodell M, Pan S, Peck J, Peterson M, Rowe W, Sanders R, Scott J, Simpson M, Smith T, Sprague A, Stockwell T, Turner R, Venter E, Wang M, Wen M, Wu D, Wu M, Xia A, Zandieh A, Zhu X. The sequence of the human genome. Science 2001; 291:1304-51. [PMID: 11181995 DOI: 10.1126/science.1058040] [Citation(s) in RCA: 7818] [Impact Index Per Article: 325.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/13/2022]
Abstract
A 2.91-billion base pair (bp) consensus sequence of the euchromatic portion of the human genome was generated by the whole-genome shotgun sequencing method. The 14.8-billion bp DNA sequence was generated over 9 months from 27,271,853 high-quality sequence reads (5.11-fold coverage of the genome) from both ends of plasmid clones made from the DNA of five individuals. Two assembly strategies-a whole-genome assembly and a regional chromosome assembly-were used, each combining sequence data from Celera and the publicly funded genome effort. The public data were shredded into 550-bp segments to create a 2.9-fold coverage of those genome regions that had been sequenced, without including biases inherent in the cloning and assembly procedure used by the publicly funded group. This brought the effective coverage in the assemblies to eightfold, reducing the number and size of gaps in the final assembly over what would be obtained with 5.11-fold coverage. The two assembly strategies yielded very similar results that largely agree with independent mapping data. The assemblies effectively cover the euchromatic regions of the human chromosomes. More than 90% of the genome is in scaffold assemblies of 100,000 bp or more, and 25% of the genome is in scaffolds of 10 million bp or larger. Analysis of the genome sequence revealed 26,588 protein-encoding transcripts for which there was strong corroborating evidence and an additional approximately 12,000 computationally derived genes with mouse matches or other weak supporting evidence. Although gene-dense clusters are obvious, almost half the genes are dispersed in low G+C sequence separated by large tracts of apparently noncoding sequence. Only 1.1% of the genome is spanned by exons, whereas 24% is in introns, with 75% of the genome being intergenic DNA. Duplications of segmental blocks, ranging in size up to chromosomal lengths, are abundant throughout the genome and reveal a complex evolutionary history. Comparative genomic analysis indicates vertebrate expansions of genes associated with neuronal function, with tissue-specific developmental regulation, and with the hemostasis and immune systems. DNA sequence comparisons between the consensus sequence and publicly funded genome data provided locations of 2.1 million single-nucleotide polymorphisms (SNPs). A random pair of human haploid genomes differed at a rate of 1 bp per 1250 on average, but there was marked heterogeneity in the level of polymorphism across the genome. Less than 1% of all SNPs resulted in variation in proteins, but the task of determining which SNPs have functional consequences remains an open challenge.
Collapse
|
|
24 |
7818 |
2
|
Adams MD, Celniker SE, Holt RA, Evans CA, Gocayne JD, Amanatides PG, Scherer SE, Li PW, Hoskins RA, Galle RF, George RA, Lewis SE, Richards S, Ashburner M, Henderson SN, Sutton GG, Wortman JR, Yandell MD, Zhang Q, Chen LX, Brandon RC, Rogers YH, Blazej RG, Champe M, Pfeiffer BD, Wan KH, Doyle C, Baxter EG, Helt G, Nelson CR, Gabor GL, Abril JF, Agbayani A, An HJ, Andrews-Pfannkoch C, Baldwin D, Ballew RM, Basu A, Baxendale J, Bayraktaroglu L, Beasley EM, Beeson KY, Benos PV, Berman BP, Bhandari D, Bolshakov S, Borkova D, Botchan MR, Bouck J, Brokstein P, Brottier P, Burtis KC, Busam DA, Butler H, Cadieu E, Center A, Chandra I, Cherry JM, Cawley S, Dahlke C, Davenport LB, Davies P, de Pablos B, Delcher A, Deng Z, Mays AD, Dew I, Dietz SM, Dodson K, Doup LE, Downes M, Dugan-Rocha S, Dunkov BC, Dunn P, Durbin KJ, Evangelista CC, Ferraz C, Ferriera S, Fleischmann W, Fosler C, Gabrielian AE, Garg NS, Gelbart WM, Glasser K, Glodek A, Gong F, Gorrell JH, Gu Z, Guan P, Harris M, Harris NL, Harvey D, Heiman TJ, Hernandez JR, Houck J, Hostin D, Houston KA, Howland TJ, Wei MH, Ibegwam C, Jalali M, Kalush F, Karpen GH, Ke Z, Kennison JA, Ketchum KA, Kimmel BE, Kodira CD, Kraft C, Kravitz S, Kulp D, Lai Z, Lasko P, Lei Y, Levitsky AA, Li J, Li Z, Liang Y, Lin X, Liu X, Mattei B, McIntosh TC, McLeod MP, McPherson D, Merkulov G, Milshina NV, Mobarry C, Morris J, Moshrefi A, Mount SM, Moy M, Murphy B, Murphy L, Muzny DM, Nelson DL, Nelson DR, Nelson KA, Nixon K, Nusskern DR, Pacleb JM, Palazzolo M, Pittman GS, Pan S, Pollard J, Puri V, Reese MG, Reinert K, Remington K, Saunders RD, Scheeler F, Shen H, Shue BC, Sidén-Kiamos I, Simpson M, Skupski MP, Smith T, Spier E, Spradling AC, Stapleton M, Strong R, Sun E, Svirskas R, Tector C, Turner R, Venter E, Wang AH, Wang X, Wang ZY, Wassarman DA, Weinstock GM, Weissenbach J, Williams SM, Worley KC, Wu D, Yang S, Yao QA, Ye J, Yeh RF, Zaveri JS, Zhan M, Zhang G, Zhao Q, Zheng L, Zheng XH, Zhong FN, Zhong W, Zhou X, Zhu S, Zhu X, Smith HO, Gibbs RA, Myers EW, Rubin GM, Venter JC. The genome sequence of Drosophila melanogaster. Science 2000; 287:2185-95. [PMID: 10731132 DOI: 10.1126/science.287.5461.2185] [Citation(s) in RCA: 4043] [Impact Index Per Article: 161.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/02/2022]
Abstract
The fly Drosophila melanogaster is one of the most intensively studied organisms in biology and serves as a model system for the investigation of many developmental and cellular processes common to higher eukaryotes, including humans. We have determined the nucleotide sequence of nearly all of the approximately 120-megabase euchromatic portion of the Drosophila genome using a whole-genome shotgun sequencing strategy supported by extensive clone-based sequence and a high-quality bacterial artificial chromosome physical map. Efforts are under way to close the remaining gaps; however, the sequence is of sufficient accuracy and contiguity to be declared substantially complete and to support an initial analysis of genome structure and preliminary gene annotation and interpretation. The genome encodes approximately 13,600 genes, somewhat fewer than the smaller Caenorhabditis elegans genome, but with comparable functional diversity.
Collapse
|
|
25 |
4043 |
3
|
Myers EW, Sutton GG, Delcher AL, Dew IM, Fasulo DP, Flanigan MJ, Kravitz SA, Mobarry CM, Reinert KH, Remington KA, Anson EL, Bolanos RA, Chou HH, Jordan CM, Halpern AL, Lonardi S, Beasley EM, Brandon RC, Chen L, Dunn PJ, Lai Z, Liang Y, Nusskern DR, Zhan M, Zhang Q, Zheng X, Rubin GM, Adams MD, Venter JC. A whole-genome assembly of Drosophila. Science 2000; 287:2196-204. [PMID: 10731133 DOI: 10.1126/science.287.5461.2196] [Citation(s) in RCA: 997] [Impact Index Per Article: 39.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/02/2022]
Abstract
We report on the quality of a whole-genome assembly of Drosophila melanogaster and the nature of the computer algorithms that accomplished it. Three independent external data sources essentially agree with and support the assembly's sequence and ordering of contigs across the euchromatic portion of the genome. In addition, there are isolated contigs that we believe represent nonrepetitive pockets within the heterochromatin of the centromeres. Comparison with a previously sequenced 2.9- megabase region indicates that sequencing accuracy within nonrepetitive segments is greater than 99. 99% without manual curation. As such, this initial reconstruction of the Drosophila sequence should be of substantial value to the scientific community.
Collapse
|
|
25 |
997 |
4
|
Lai Z, Colón W, Kelly JW. The acid-mediated denaturation pathway of transthyretin yields a conformational intermediate that can self-assemble into amyloid. Biochemistry 1996; 35:6470-82. [PMID: 8639594 DOI: 10.1021/bi952501g] [Citation(s) in RCA: 449] [Impact Index Per Article: 15.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023]
Abstract
Transthyretin (TTR) amyloid fibril formation is observed during partial acid denaturation and while refolding acid-denatured TTR, implying that amyloid fibril formation results from the self-assembly of a conformational intermediate. The acid denaturation pathway of TTR has been studied in detail herein employing a variety of biophysical methods to characterize the intermediate(s) capable of amyloid fibril formation. At physiological concentrations, tetrameric TTR remains associated from pH 7 to pH 5 and is incapable of amyloid fibril formation. Tetrameric TTR dissociates to a monomer in a process that is dependent on both pH and protein concentration below pH 5. The extent of amyloid fibril formation correlates with the concentration of the TTR monomer having an altered, but defined, tertiary structure over the pH range of 5.0-3.9. The inherent Trp fluorescence-monitored denaturation curve of TTR exhibits a plateau over the pH range where amyloid fibril formation is observed (albeit at a higher concentration), implying that a steady-state concentration of the amyloidogenic intermediate with an altered tertiary structure is being detected. Interestingly, 1-anilino-8-naphthalenesulfonate fluorescence is at a minimum at the pH associated with maximal amyloid fibril formation (pH 4.4), implying that the amyloidogenic intermediate does not have a high extent of hydrophobic surface area exposed, consistent with a defined tertiary structure. Transthyretin has two Trp residues in its primary structure, Trp-41 and Trp-79, which are conveniently located far apart in the tertiary structure of TTR. Replacement of each Trp with Phe affords two single Trp containing variants which were used to probe local pH-dependent tertiary structural changes proximal to these chromophores. The pH-dependent fluorescence behavior of the Trp-79-Phe mutant strongly suggests that Trp-41 is located near the site of the tertiary structural rearrangement that occurs in the formation of the monomeric amyloidogenic intermediate, likely involving the C-strand-loop-D-strand region. Upon further acidification of TTR (below pH 4.4), the structurally defined monomeric amyloidogenic intermediate begins to adopt alternative conformations that are not amyloidogenic, ultimately forming an A-state conformation below pH 3 which is also not amyloidogenic. In summary, analytical equilibrium ultracentrifugation, SDS-PAGE, far- and near-UV CD, fluorescence, and light scattering studies suggest that the amyloidogenic intermediate is a monomeric predominantly beta-sheet structure having a well-defined tertiary structure.
Collapse
|
|
29 |
449 |
5
|
Gardner MJ, Tettelin H, Carucci DJ, Cummings LM, Aravind L, Koonin EV, Shallom S, Mason T, Yu K, Fujii C, Pederson J, Shen K, Jing J, Aston C, Lai Z, Schwartz DC, Pertea M, Salzberg S, Zhou L, Sutton GG, Clayton R, White O, Smith HO, Fraser CM, Adams MD, Venter JC, Hoffman SL. Chromosome 2 sequence of the human malaria parasite Plasmodium falciparum. Science 1998; 282:1126-32. [PMID: 9804551 DOI: 10.1126/science.282.5391.1126] [Citation(s) in RCA: 370] [Impact Index Per Article: 13.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/02/2022]
Abstract
Chromosome 2 of Plasmodium falciparum was sequenced; this sequence contains 947,103 base pairs and encodes 210 predicted genes. In comparison with the Saccharomyces cerevisiae genome, chromosome 2 has a lower gene density, introns are more frequent, and proteins are markedly enriched in nonglobular domains. A family of surface proteins, rifins, that may play a role in antigenic variation was identified. The complete sequencing of chromosome 2 has shown that sequencing of the A+T-rich P. falciparum genome is technically feasible.
Collapse
MESH Headings
- Amino Acid Sequence
- Animals
- Antigens, Protozoan/chemistry
- Antigens, Protozoan/genetics
- Base Composition
- Chromosomes/genetics
- Evolution, Molecular
- Genes, Protozoan
- Genome, Protozoan
- Introns
- Membrane Proteins/chemistry
- Membrane Proteins/genetics
- Molecular Sequence Data
- Multigene Family
- Physical Chromosome Mapping
- Plasmodium falciparum/genetics
- Protozoan Proteins/chemistry
- Protozoan Proteins/genetics
- RNA, Protozoan/genetics
- RNA, Transfer, Glu/genetics
- Repetitive Sequences, Nucleic Acid
- Reverse Transcriptase Polymerase Chain Reaction
- Sequence Alignment
- Sequence Analysis, DNA
Collapse
|
|
27 |
370 |
6
|
Bellugi U, Lichtenberger L, Jones W, Lai Z, St George M. I. The neurocognitive profile of Williams Syndrome: a complex pattern of strengths and weaknesses. J Cogn Neurosci 2000; 12 Suppl 1:7-29. [PMID: 10953231 DOI: 10.1162/089892900561959] [Citation(s) in RCA: 367] [Impact Index Per Article: 14.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/04/2022]
Abstract
The rare, genetically based disorder, Williams syndrome (WMS), produces a constellation of distinctive cognitive, neuroanatomical, and electrophysiological features which we explore through the series of studies reported here. In this paper, we focus primarily on the cognitive characteristics of WMS and begin to forge links among these characteristics, the brain, and the genetic basis of the disorder. The distinctive cognitive profile of individuals with WMS includes relative strengths in language and facial processing and profound impairment in spatial cognition. The cognitive profile of abilities, including what is 'typical' for individuals with WMS is discussed, but we also highlight areas of variability across the group of individuals with WMS that we have studied. Although the overall cognitive abilities (IQs) of individuals with WMS are typically in the mild-to-moderate range of mental retardation, the peaks and valleys within different cognitive domains make this syndrome especially intriguing to study across levels. Understanding the brain basis (and ultimately the genetic basis) for higher cognitive functioning is the goal we have begun to undertake with this line of interdisciplinary research.
Collapse
|
Review |
25 |
367 |
7
|
Cruz C, Castroviejo-Bermejo M, Gutiérrez-Enríquez S, Llop-Guevara A, Ibrahim YH, Gris-Oliver A, Bonache S, Morancho B, Bruna A, Rueda OM, Lai Z, Polanska UM, Jones GN, Kristel P, de Bustos L, Guzman M, Rodríguez O, Grueso J, Montalban G, Caratú G, Mancuso F, Fasani R, Jiménez J, Howat WJ, Dougherty B, Vivancos A, Nuciforo P, Serres-Créixams X, Rubio IT, Oaknin A, Cadogan E, Barrett JC, Caldas C, Baselga J, Saura C, Cortés J, Arribas J, Jonkers J, Díez O, O'Connor MJ, Balmaña J, Serra V. RAD51 foci as a functional biomarker of homologous recombination repair and PARP inhibitor resistance in germline BRCA-mutated breast cancer. Ann Oncol 2018; 29:1203-1210. [PMID: 29635390 PMCID: PMC5961353 DOI: 10.1093/annonc/mdy099] [Citation(s) in RCA: 307] [Impact Index Per Article: 43.9] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/24/2022] Open
Abstract
Background BRCA1 and BRCA2 (BRCA1/2)-deficient tumors display impaired homologous recombination repair (HRR) and enhanced sensitivity to DNA damaging agents or to poly(ADP-ribose) polymerase (PARP) inhibitors (PARPi). Their efficacy in germline BRCA1/2 (gBRCA1/2)-mutated metastatic breast cancers has been recently confirmed in clinical trials. Numerous mechanisms of PARPi resistance have been described, whose clinical relevance in gBRCA-mutated breast cancer is unknown. This highlights the need to identify functional biomarkers to better predict PARPi sensitivity. Patients and methods We investigated the in vivo mechanisms of PARPi resistance in gBRCA1 patient-derived tumor xenografts (PDXs) exhibiting differential response to PARPi. Analysis included exome sequencing and immunostaining of DNA damage response proteins to functionally evaluate HRR. Findings were validated in a retrospective sample set from gBRCA1/2-cancer patients treated with PARPi. Results RAD51 nuclear foci, a surrogate marker of HRR functionality, were the only common feature in PDX and patient samples with primary or acquired PARPi resistance. Consistently, low RAD51 was associated with objective response to PARPi. Evaluation of the RAD51 biomarker in untreated tumors was feasible due to endogenous DNA damage. In PARPi-resistant gBRCA1 PDXs, genetic analysis found no in-frame secondary mutations, but BRCA1 hypomorphic proteins in 60% of the models, TP53BP1-loss in 20% and RAD51-amplification in one sample, none mutually exclusive. Conversely, one of three PARPi-resistant gBRCA2 tumors displayed BRCA2 restoration by exome sequencing. In PDXs, PARPi resistance could be reverted upon combination of a PARPi with an ataxia-telangiectasia mutated (ATM) inhibitor. Conclusion Detection of RAD51 foci in gBRCA tumors correlates with PARPi resistance regardless of the underlying mechanism restoring HRR function. This is a promising biomarker to be used in the clinic to better select patients for PARPi therapy. Our study also supports the clinical development of PARPi combinations such as those with ATM inhibitors.
Collapse
|
Research Support, N.I.H., Extramural |
7 |
307 |
8
|
Jones W, Bellugi U, Lai Z, Chiles M, Reilly J, Lincoln A, Adolphs R. II. Hypersociability in Williams Syndrome. J Cogn Neurosci 2000; 12 Suppl 1:30-46. [PMID: 10953232 DOI: 10.1162/089892900561968] [Citation(s) in RCA: 298] [Impact Index Per Article: 11.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/04/2022]
Abstract
Studies of abnormal populations provide a rare opportunity for examining relationships between cognition, genotype and brain neurobiology, permitting comparisons across these different levels of analysis. In our studies, we investigate individuals with a rare, genetically based disorder called Williams syndrome (WMS) to draw links among these levels. A critical component of such a cross-domain undertaking is the clear delineation of the phenotype of the disorder in question. Of special interest in this paper is a relatively unexplored unusual social phenotype in WMS that includes an overfriendly and engaging personality. Four studies measuring distinct aspects of hypersocial behavior in WMS are presented, each probing specific aspects in WMS infants, toddlers, school age children, and adults. The abnormal profile of excessively social behavior represents an important component of the phenotype that may distinguish WMS from other developmental disorders. Furthermore, the studies show that the profile is observed across a wide range of ages, and emerges consistently across multiple experimental paradigms. These studies of hypersocial behavior in WMS promise to provide the groundwork for crossdisciplinary analyses of gene-brain-behavior relationships.
Collapse
|
|
25 |
298 |
9
|
Miroy GJ, Lai Z, Lashuel HA, Peterson SA, Strang C, Kelly JW. Inhibiting transthyretin amyloid fibril formation via protein stabilization. Proc Natl Acad Sci U S A 1996; 93:15051-6. [PMID: 8986762 PMCID: PMC26354 DOI: 10.1073/pnas.93.26.15051] [Citation(s) in RCA: 256] [Impact Index Per Article: 8.8] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 06/17/1996] [Accepted: 09/30/1996] [Indexed: 02/03/2023] Open
Abstract
Transthyretin (TTR) amyloid fibril formation is observed systemically in familial amyloid polyneuropathy and senile systemic amyloidosis and appears to be the causative agent in these diseases. Herein, we demonstrate conclusively that thyroxine (10.8 microM) inhibits TTR fibril formation efficiently in vitro and does so by stabilizing the tetramer against dissociation and the subsequent conformational changes required for amyloid fibril formation. In addition, the nonnative ligand 2,4,6-triiodophenol, which binds to TTR with slightly increased affinity also inhibits TTR fibril formation by this mechanism. Sedimentation velocity experiments were employed to show that TTR undergoes dissociation (linked to a conformational change) to form the monomeric amyloidogenic intermediate, which self-assembles into amyloid in the absence, but not in the presence of thyroxine. These results demonstrate the feasibility of using small molecules to stabilize the native fold of a potentially amyloidogenic human protein, thus preventing the conformational changes, which appear to be the common link in several human amyloid diseases. This strategy and the compounds resulting from further development should prove useful for critically evaluating the amyloid hypothesis--i.e., the putative cause-and-effect relationship between TTR amyloid deposition and the onset of familial amyloid polyneuropathy and senile systemic amyloidosis.
Collapse
|
research-article |
29 |
256 |
10
|
Lai Z. Kinetic growth with surface relaxation: Continuum versus atomistic models. PHYSICAL REVIEW LETTERS 1991; 66:2348-2351. [PMID: 10043462 DOI: 10.1103/physrevlett.66.2348] [Citation(s) in RCA: 230] [Impact Index Per Article: 6.8] [Reference Citation Analysis] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 05/23/2023]
|
|
34 |
230 |
11
|
Reiss AL, Eliez S, Schmitt JE, Straus E, Lai Z, Jones W, Bellugi U. IV. Neuroanatomy of Williams syndrome: a high-resolution MRI study. J Cogn Neurosci 2000; 12 Suppl 1:65-73. [PMID: 10953234 DOI: 10.1162/089892900561986] [Citation(s) in RCA: 179] [Impact Index Per Article: 7.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/04/2022]
Abstract
Williams syndrome (WMS), a genetic condition resulting from a contiguous deletion on the long arm of chromosome 7, is associated with a relatively consistent profile of neurocognitive and neurobehavioral features. The distinctiveness and regularity of the profile of learning and behavioral characteristics in this genetic condition suggests that underlying neurobiological correlates may be identifiable. In this initial study, we report findings derived from a high-resolution neuroimaging study of 14 young adult subjects with WMS and an individually matched normal control group. Compared to controls, subjects with WMS were noted to have decreased overall brain and cerebral volumes, relative preservation of cerebellar and superior temporal gyrus (STG) volumes, and disproportionate volume reduction of the brainstem. Analyses also suggested that the pattern of cerebral lobe proportions in WMS may be altered compared to normal controls with a greater ratio of frontal to posterior (parietal+occipital) tissue. Assessment of tissue composition indicated that, relative to controls, individuals with WMS have relative preservation of cerebral gray matter volume and disproportionate reduction in cerebral white matter volume. However, within the cerebral gray matter tissue compartment, the right occipital lobe was noted to have excess volume loss. Combined with our growing knowledge of the function of genes in the commonly deleted region for WMS, more detailed information regarding the structure and function of the WMS brain will provide a unique opportunity for elucidating meaningful correlations amongst genetic, neurobiological, and neurobehavioral factors in humans.
Collapse
|
Clinical Trial |
25 |
179 |
12
|
Korenberg JR, Chen XN, Hirota H, Lai Z, Bellugi U, Burian D, Roe B, Matsuoka R. VI. Genome structure and cognitive map of Williams syndrome. J Cogn Neurosci 2000; 12 Suppl 1:89-107. [PMID: 10953236 DOI: 10.1162/089892900562002] [Citation(s) in RCA: 171] [Impact Index Per Article: 6.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/04/2022]
Abstract
Williams syndrome (WMS) is a most compelling model of human cognition, of human genome organization, and of evolution. Due to a deletion in chromosome band 7q11.23, subjects have cardiovascular, connective tissue, and neurodevelopmental deficits. Given the striking peaks and valleys in neurocognition including deficits in visual-spatial and global processing, preserved language and face processing, hypersociability, and heightened affect, the goal of this work has been to identify the genes that are responsible, the cause of the deletion, and its origin in primate evolution. To do this, we have generated an integrated physical, genetic, and transcriptional map of the WMS and flanking regions using multicolor metaphase and interphase fluorescence in situ hybridization (FISH) of bacterial artificial chromosomes (BACs) and P1 artificial chromosomes (PACs), BAC end sequencing, PCR gene marker and microsatellite, large-scale sequencing, cDNA library, and database analyses. The results indicate the genomic organization of the WMS region as two nested duplicated regions flanking a largely single-copy region. There are at least two common deletion breakpoints, one in the centromeric and at least two in the telomeric repeated regions. Clones anchoring the unique to the repeated regions are defined along with three new pseudogene families. Primate studies indicate an evolutionary hot spot for chromosomal inversion in the WMS region. A cognitive phenotypic map of WMS is presented, which combines previous data with five further WMS subjects and three atypical WMS subjects with deletions; two larger (deleted for D7S489L) and one smaller, deleted for genes telomeric to FZD9, through LIMK1, but not WSCR1 or telomeric. The results establish regions and consequent gene candidates for WMS features including mental retardation, hypersociability, and facial features. The approach provides the basis for defining pathways linking genetic underpinnings with the neuroanatomical, functional, and behavioral consequences that result in human cognition.
Collapse
|
Clinical Trial |
25 |
171 |
13
|
Lashuel HA, Lai Z, Kelly JW. Characterization of the transthyretin acid denaturation pathways by analytical ultracentrifugation: implications for wild-type, V30M, and L55P amyloid fibril formation. Biochemistry 1998; 37:17851-64. [PMID: 9922152 DOI: 10.1021/bi981876+] [Citation(s) in RCA: 159] [Impact Index Per Article: 5.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
Abstract
Analytical ultracentrifugation methods were utilized to further characterize the acid denaturation pathways of wild-type, V30M, and L55P transthyretin (TTR) that generate intermediates leading to amyloid fibril formation and possibly the diseases senile systemic amyloidosis and familial amyloid polyneuropathy. Equilibrium and velocity methods were employed herein to characterize the TTR quaternary structural requirements for amyloid fibril formation. From neutral to slightly acidic conditions (pH 7.5-5.1), wild-type transthyretin (0.2-0.3 mg/mL, 100 mM KCl, 37 degrees C) exists as a tetramer and is incapable of fibril formation. Under more acidic conditions (pH 5 to 3.9), tetrameric wild-type TTR slowly dissociates to a monomer having an alternatively folded tertiary structure(s) that self-assembles at physiological concentration (0.2 mg/mL) into a ladder of quaternary structural intermediates of increasing molecular weight. These intermediates appear to be on the pathway of amyloid fibril formation, since they ultimately disappear when amyloid fibrils are observed. The V30M and L55P TTR variants exhibit similar acid denaturation pathways, with the exception that dissociation of the tetramer to the monomeric amyloidogenic intermediate occurs at a higher pH and to a much greater extent, allowing the quaternary structural intermediates to be readily observed by velocity methods. Partial denaturation and assembly of the monomeric amyloidogenic intermediate(s) occur at pH 5.4 for V30M and L55P TTR over a 72 h period, during which wild-type TTR maintains its normal tetrameric three-dimensional structure. Interestingly, the L55P and V30M familial amyloid polyneuropathy (FAP) associated variants form amyloid protofilaments at pH 7.5 (37 degrees C) after several weeks of incubation, suggesting that the activation barriers for TTR tetramer dissociation to the monomeric amyloidogenic intermediate are much lower for the FAP variants relative to wild-type TTR, which does not form amyloid or amyloid protofilaments under these conditions. This study establishes the key role of the monomeric amyloidogenic intermediate and its self-assembly into a ladder of quaternary structural intermediates for the formation of wild-type, V30M, and L55P transthyretin amyloid fibrils.
Collapse
|
|
27 |
159 |
14
|
McCutchen SL, Lai Z, Miroy GJ, Kelly JW, Colón W. Comparison of lethal and nonlethal transthyretin variants and their relationship to amyloid disease. Biochemistry 1995; 34:13527-36. [PMID: 7577941 DOI: 10.1021/bi00041a032] [Citation(s) in RCA: 152] [Impact Index Per Article: 5.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/26/2023]
Abstract
The role that transthyretin (TTR) mutations play in the amyloid disease familial amyloid polyneuropathy (FAP) has been probed by comparing the biophysical properties of several TTR variants as a function of pH. We have previously demonstrated that the partial acid denaturation of TTR is sufficient to effect amyloid fibril formation by self-assembly of a denaturation intermediate which appears to be monomeric. Earlier studies on the most pathogenic FAP variant known, Leu-55-Pro, revealed that this variant is much less stable toward acid denaturation than wild-type TTR, apparently explaining why this variant can form amyloid fibrils under mildly acidic conditions where wild-type TTR remains nonamyloidogenic. The hypothesis that FAP mutations destabilize the TTR tetramer in favor of a monomeric amyloidogenic intermediate under lysosomal (acidic) conditions is further supported by the data described here. We compare the acid stability and amyloidogenicity of the most prevalent FAP variant, Val-30-Met, along with the double mutant, Val-30-Met/Thr-119-Met, which serves to model the effects of these mutations in heterozygous patients where the mutations are in different subunits. In addition, we have characterized the Thr-119-Met TTR variant, which is a common nonpathogenic variant in the Portuguese population, to further investigate the role that this mutation plays in protecting individuals who also carry the Val-30-Met mutation against the classically severe FAP pathology. This biophysical study demonstrates that Val-30-Met TTR is significantly less stable toward acid denaturation and more amyloidogenic than wild-type TTR, which in turn is less stable and more amyloidogenic than Thr-119-Met TTR. Interestingly, the double mutant Val-30-Met/Thr-119-Met is very similar to wild-type TTR in terms of its stability toward acid denaturation and its amyloidogenicity. The data suggest that the Thr-119-Met mutation confers decreased amyloidogenicity by stabilizing tetrameric TTR toward acid denaturation. In addition, fluorescence studies monitoring the acid-mediated denaturation pathways of several TTR variants reveal that the majority exhibit a plateau in the relative fluorescence intensity over the amyloid-forming pH range, i.e., ca. pH 4.3-3.3. This intensity plateau suggests that the amyloidogenic intermediate(s) is (are) being observed over this pH range. The Thr-119-Met variant does not exhibit this plateau presumably because the amyloidogenic intermediate(s) do(es) not build up in concentration in this variant. The intermediate is undoubtedly forming in the Thr-119-Met variant, as it will form amyloid fibrils at high concentrations; however, the intermediate is only present at a low steady-state concentration which makes it difficult to detect.
Collapse
|
Comparative Study |
30 |
152 |
15
|
Nettleton EJ, Sunde M, Lai Z, Kelly JW, Dobson CM, Robinson CV. Protein subunit interactions and structural integrity of amyloidogenic transthyretins: evidence from electrospray mass spectrometry. J Mol Biol 1998; 281:553-64. [PMID: 9698569 DOI: 10.1006/jmbi.1998.1937] [Citation(s) in RCA: 148] [Impact Index Per Article: 5.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
Wild-type and variant transthyretins form amyloid fibrils in two different diseases. The biologically active form of transthyretin is a tetramer but there is evidence that a monomeric species is the amyloidogenic intermediate. Using mass spectrometry we have developed an approach to monitor the proportions of monomer and tetramer in wild-type and variant transthyretins, and found a strong correlation between the instability of the tetramer in the gas phase and the amyloidogenicity of the protein variant. The presence of water molecules in the central channel has been found to be critical for maintaining intact the complex in the gas phase, with additional stability observed in the presence of excess thyroxine. The solution structure of monomeric transthyretin under fibril-forming conditions was studied using hydrogen exchange monitored by mass spectrometry. The results show that Val30Met transthyretin, the commonest amyloidogenic variant, exhibits loss of hydrogen exchange protection substantially more rapidly than the wild-type protein, suggesting partial unfolding of the beta-sheet structure. These results provide new insights into the correlation between tetramer stability and amyloidogenicity as well as supporting a possible route to fibril formation via transient unfolding of the transthyretin monomer.
Collapse
|
|
27 |
148 |
16
|
Tian M, Broxmeyer HE, Fan Y, Lai Z, Zhang S, Aronica S, Cooper S, Bigsby RM, Steinmetz R, Engle SJ, Mestek A, Pollock JD, Lehman MN, Jansen HT, Ying M, Stambrook PJ, Tischfield JA, Yu L. Altered hematopoiesis, behavior, and sexual function in mu opioid receptor-deficient mice. J Exp Med 1997; 185:1517-22. [PMID: 9126934 PMCID: PMC2196276 DOI: 10.1084/jem.185.8.1517] [Citation(s) in RCA: 139] [Impact Index Per Article: 5.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 02/10/1997] [Indexed: 02/04/2023] Open
Abstract
The mu opioid receptor is thought to be the cellular target of opioid narcotics such as morphine and heroin, mediating their effects in both pain relief and euphoria. Its involvement is also implicated in a range of diverse biological processes. Using a mouse model in which the receptor gene was disrupted by targeted homologous recombination, we explored the involvement of this receptor in a number of physiological functions. Mice homozygous for the disrupted gene developed normally, but their motor function was altered. Drug-naive homozygotes displayed reduced locomotor activity, and morphine did not induce changes in locomotor activity observed in wild-type mice. Unexpectedly, lack of a functional receptor resulted in changes in both the host defense system and the reproductive system. We observed increased proliferation of granulocyte-macrophage, erythroid, and multipotential progenitor cells in both bone marrow and spleen, indicating a link between hematopoiesis and the opioid system, both of which are stress-responsive systems. Unexpected changes in sexual function in male homozygotes were also observed, as shown by reduced mating activity, a decrease in sperm count and motility, and smaller litter size. Taken together, these results suggest a novel role of the mu opioid receptor in hematopoiesis and reproductive physiology, in addition to its known involvement in pain relief.
Collapse
|
research-article |
28 |
139 |
17
|
Morganstein DL, Lai Z, Spain L, Diem S, Levine D, Mace C, Gore M, Larkin J. Thyroid abnormalities following the use of cytotoxic T-lymphocyte antigen-4 and programmed death receptor protein-1 inhibitors in the treatment of melanoma. Clin Endocrinol (Oxf) 2017; 86:614-620. [PMID: 28028828 DOI: 10.1111/cen.13297] [Citation(s) in RCA: 137] [Impact Index Per Article: 17.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Submit a Manuscript] [Subscribe] [Scholar Register] [Received: 08/15/2016] [Revised: 11/14/2016] [Accepted: 12/20/2016] [Indexed: 01/21/2023]
Abstract
CONTEXT Checkpoint inhibitors are emerging as important cancer therapies but are associated with a high rate of immune side effects, including endocrinopathy. OBJECTIVE To determine the burden of thyroid dysfunction in patients with melanoma treated with immune checkpoint inhibitors and describe the clinical course. DESIGN AND PATIENTS Consecutive patients with melanoma treated with either ipilimumab, nivolumab, pembrolizumab or the combination of ipilimumab and nivolumab were identified. Baseline thyroid function tests were used to exclude those with pre-existing thyroid abnormalities, and thyroid function tests during treatment used to identify those with thyroid dysfunction. RESULTS Rates of overt thyroid dysfunction were in keeping with the published phase 3 trials. Hypothyroidism occurred in 13·0% treated with a programmed death receptor-1 (PD-1) inhibitor and 22·2% with a combination of PD-1 inhibitor and ipilimumab. Transient subclinical hyperthyroidism was observed in 13·0% treated with a PD-1 inhibitor, 15·9% following a PD-1 inhibitor, and 22·2% following combination treatment with investigations suggesting a thyroiditic mechanism rather than Graves' disease, and a high frequency of subsequent hypothyroidism. Any thyroid abnormality occurred in 23·0% following ipilimumab, 39·1% following a PD-1 inhibitor and 50% following combination treatment. Abnormal thyroid function was more common in female patients. CONCLUSION Thyroid dysfunction occurs commonly in patients with melanoma treated with immune checkpoint inhibitors, with rates, including subclinical dysfunction, occurring in up to 50%.
Collapse
|
|
8 |
137 |
18
|
Kelly JW, Colon W, Lai Z, Lashuel HA, McCulloch J, McCutchen SL, Miroy GJ, Peterson SA. Transthyretin quaternary and tertiary structural changes facilitate misassembly into amyloid. ADVANCES IN PROTEIN CHEMISTRY 1997; 50:161-81. [PMID: 9338081 DOI: 10.1016/s0065-3233(08)60321-6] [Citation(s) in RCA: 131] [Impact Index Per Article: 4.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
Human transthyretin (TTR) can be transformed into amyloid fibrils by partial acid denaturation to yield a monomeric amyloidogenic intermediate that self-associates into amyloid through quaternary structural intermediates, which are identified by sedimentation velocity methods. The monomeric amyloidogenic intermediate has substantial beta-sheet structure with a nonnative but intact tertiary structure as discerned from spectroscopic methods. Proteolysis sensitivity studies suggest that the C-strand-loop-D-strand portion of TTR becomes disordered and moves away from the core of the beta-sandwich fold upon formation of the monomeric amyloidogenic intermediate over the pH range 5.1-3.9. The single site mutations that are associated with early onset amyloid disease [familial amyloid polyneuropathy (FAP)] function by destabilizing tetrameric TTR. Under mild denaturing conditions, the FAP variants populate the monomeric amyloidogenic intermediate conformation, which assembles into amyloid, whereas wild-type TTR remains tetrameric and nonamyloidogenic. The FAP mutations do not significantly alter the native folded structure; instead, they appear to act by making the thermodynamics and perhaps the kinetics more favorable for formation of the amyloidogenic intermediate. Suppressor mutations have also been characterized that strongly stabilize tetrameric TTR and disfavor the formation of the monomeric amyloidogenic intermediate, thus inhibiting amyloid formation. The mechanistic details characterizing transthyretin amyloid fibril formation available from the biophysical studies outlined within have been utilized to develop a new therapeutic strategy for intervention in human amyloid disease. This approach features small molecules that bind with high affinity to the normal fold of transthyretin, inhibiting the quaternary and tertiary structural changes associated with the formation of the monomeric amyloidogenic intermediate that self-assembles into amyloid. Ligand binding to TTR stabilizes the native tetrameric fold, which is nonamyloidogenic.
Collapse
|
Review |
28 |
131 |
19
|
Lai Z, Ferry KV, Diamond MA, Wee KE, Kim YB, Ma J, Yang T, Benfield PA, Copeland RA, Auger KR. Human mdm2 mediates multiple mono-ubiquitination of p53 by a mechanism requiring enzyme isomerization. J Biol Chem 2001; 276:31357-67. [PMID: 11397792 DOI: 10.1074/jbc.m011517200] [Citation(s) in RCA: 117] [Impact Index Per Article: 4.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023] Open
Abstract
The mdm2 gene product is an important regulator of p53 function and stability. mdm2 is an E3 ubiquitin ligase for p53 and the RING finger domain of mdm2 is critical for ligase activity. Ubiquitin (Ub) conjugation is a general targeting modification and poly-ubiquitin chains specifically target proteins to the proteasome for degradation. In this report, we show that the multistep cascade of mdm2-mediated p53 ubiquitination can be reduced to three purified recombinant proteins: ubiquitin-conjugated E2, mdm2, and p53. This simplification allows enzymatic analysis of the isolated ligase reaction. The simplified reaction recapitulates the ubiquitination of p53 observed with individual components and the p53-Ub((n)) is qualitatively similar to p53-Ub((n)) detected in lactacystin-treated cells. Surprisingly, we find that p53 is modified with multiple mono-ubiquitin moieties as opposed to a poly-ubiquitin chain. Finally, kinetic analysis indicates the transfer reaction proceeds either through a modified Ping Pong mechanism involving requisite enzyme isomerization steps, or through a Rapid Equilibrium Random Bi Bi mechanism involving very large anti-cooperative interactions between the two substrate binding pockets on the enzyme, mediated through allosteric changes in enzyme structure.
Collapse
|
|
24 |
117 |
20
|
Hanover JA, Lai Z, Lee G, Lubas WA, Sato SM. Elevated O-linked N-acetylglucosamine metabolism in pancreatic beta-cells. Arch Biochem Biophys 1999; 362:38-45. [PMID: 9917327 DOI: 10.1006/abbi.1998.1016] [Citation(s) in RCA: 103] [Impact Index Per Article: 4.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
High intracellular glucose concentrations increase flux though the hexosamine biosynthetic pathway, resulting in elevated UDP-N-acetylglucosamine (GlcNAc) concentrations. The nucleocytoplasmic enzyme O-linked N-acetylglucosaminyltransferase (OGT) uses UDP-GlcNAc as a donor to modify numerous critical substrates, including nuclear pore proteins and transcription factors. Here, we document (a) the overwhelming enrichment of pancreatic OGT transcripts in the beta-cells of the islets of Langerhans, (b) the physiologically significant increase in the level of O-GlcNAc residues present in beta-cells, and (c) the action of streptozotocin, a close analogue of GlcNAc, to selectively inhibit O-GlcNAcase, an enzyme involved in the removal of O-GlcNAc residues. Taken together, these findings suggest that pancreatic beta cells maintain a highly elevated O-GlcNAc metabolism and that the diabetes inducing drug streptozotocin inhibits O-GlcNAcase.
Collapse
|
|
26 |
103 |
21
|
Lai Z, Roos P, Zhai O, Olsson Y, Fhölenhag K, Larsson C, Nyberg F. Age-related reduction of human growth hormone-binding sites in the human brain. Brain Res 1993; 621:260-6. [PMID: 8242339 DOI: 10.1016/0006-8993(93)90114-3] [Citation(s) in RCA: 99] [Impact Index Per Article: 3.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/29/2023]
Abstract
Previous studies have shown that alterations in various neuroendocrine functions occur with increasing age. We here report a study of growth hormone (GH)-binding sites in different areas of post-mortem human brains collected from individual males and females of different age. The results indicate that there exists a significant negative correlation between the density of GH-binding sites and increasing age. This phenomenon was observed in both sexes in brain areas such as choroid plexus, hippocampus, hypothalamus, pituitary and putamen but not in e.g. thalamus. In all tissues (except for choroid plexus), the GH binding was significantly higher in those originating from females than those from males. This discrepancy was found likely to be associated with the affinity of GH to lactogenic rather than to somatogenic sites as no pronounced sex difference in binding was observed in the presence of excessive amounts of human prolactin. Data also indicate that the putative GH receptors in the various brain regions differ with regard to binding constants and to the estimated molecular size of the hormone-binding units. The loss of GH receptors in brain of elderly people may have consequences in several physiological courses. The decrease in GH binding at hypothalamic and pituitary levels may be of importance for the mechanisms behind the release or secretion of the hormone.
Collapse
|
|
32 |
99 |
22
|
Abstract
Previously we described safe and efficient three-component human immunodeficiency virus type 1 (HIV-1)-based gene transfer systems for delivery of genes into nondividing cells (H. Mochizuki, J. P. Schwartz, K. Tanaka, R. O. Brady, and J. Reiser, J. Virol. 72:8873-8883, 1998). To apply such vectors in anti-HIV gene therapy strategies and to express multiple proteins in single target cells, we have engineered HIV-1 vectors for the concurrent expression of multiple transgenes. Single-gene vectors, bicistronic vectors, and multigene vectors expressing up to three exogenous genes under the control of two or three different transcriptional units, placed within the viral gag-pol coding region and/or the viral nef and env genes, were designed. The genes encoding the enhanced version of green fluorescent protein (EGFP), mouse heat-stable antigen (HSA), and bacterial neomycin phosphotransferase were used as models whose expression was detected by fluorescence-activated cell sorting, fluorescence microscopy, and G418 selection. Coexpression of these reporter genes in contact-inhibited primary human skin fibroblasts (HSFs) persisted for at least 6 weeks in culture. Coexpression of the HSA and EGFP reporter genes was also achieved following cotransduction of target cells using two separate lentivirus vectors encoding HSA and EGFP, respectively. For the regulated expression of transgenes, tetracycline (Tet)-regulatable lentivirus vectors encoding the reverse Tet transactivator (rtTA) and EGFP controlled by a Tet-responsive element (TRE) were constructed. A binary HIV-1-based vector system consisting of a lentivirus encoding rtTA and a second lentivirus harboring a TRE driving the EGFP reporter gene was also designed. Doxycycline-modulated expression of the EGFP transgene was confirmed in transduced primary HSFs. These versatile vectors can potentially be used in a wide range of gene therapy applications.
Collapse
|
research-article |
25 |
92 |
23
|
Lai Z, McCulloch J, Lashuel HA, Kelly JW. Guanidine hydrochloride-induced denaturation and refolding of transthyretin exhibits a marked hysteresis: equilibria with high kinetic barriers. Biochemistry 1997; 36:10230-9. [PMID: 9254621 DOI: 10.1021/bi963195p] [Citation(s) in RCA: 78] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
Fluorescence and circular dichroism spectroscopy as well as analytical ultracentrifugation and glutaraldehyde cross-linking were utilized to evaluate the tertiary and quaternary structural changes occurring on the denaturation and reconstitution pathways of transthyretin (TTR) as a function of guanidine hydrochloride (GdnHCl) concentration. These results demonstrate that the GdnHCl-mediated denaturation and reconstitution of TTR is reversible. However, the lowest GdnHCl concentration that dissociates and unfolds transthyretin does not allow the unfolded monomer to refold to tetramer at a rate that is measurable. As a result, there is a striking hysteresis observed upon comparison of the GdnHCl-mediated denaturation and reconstitution transitions. The TTR tetramer does not dissociate into unfolded monomer until the denaturant concentration exceeds 4 M GdnHCl, whereas unfolded monomeric TTR (denatured in 7 M GdnHCl) does not refold and assemble into a native tetrameric structure until the GdnHCl concentration is reduced to less than 2 M. These results imply that a significant kinetic barrier intervenes between the folded tetramer and unfolded monomer in both the denaturation and reconstitution directions at pH 7. A kinetics study of the denaturation of TTR as a function of GdnHCl concentration yields a first-order rate constant for unfolding of (9.0 +/- 7.5) x 10(-11) s-1, estimated by extrapolation of the rate constants for the tetramer to unfolded monomer transition as a function of GdnHCl to 0 M GdnHCl. This rate is very slow; as a result, wild-type TTR is predicted to be kinetically stable as a tetrameric quaternary structure once formed. These results imply that the rate of TTR dissociation and partial unfolding to the monomeric amyloidogenic intermediate under denaturing conditions may play a role in transthyretin-based amyloid diseases.
Collapse
|
|
28 |
78 |
24
|
Liu J, Tolvgard A, Malmodin J, Lai Z. A reliable and environmentally friendly packaging technology-flip-chip joining using anisotropically conductive adhesive. ACTA ACUST UNITED AC 1999. [DOI: 10.1109/6144.774729] [Citation(s) in RCA: 77] [Impact Index Per Article: 3.0] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/09/2022]
|
|
26 |
77 |
25
|
Schmidt B, Lindman S, Tong W, Lindeberg G, Gogoll A, Lai Z, Thörnwall M, Synnergren B, Nilsson A, Welch CJ, Sohtell M, Westerlund C, Nyberg F, Karlén A, Hallberg A. Design, synthesis, and biological activities of four angiotensin II receptor ligands with gamma-turn mimetics replacing amino acid residues 3-5. J Med Chem 1997; 40:903-19. [PMID: 9083479 DOI: 10.1021/jm960553d] [Citation(s) in RCA: 68] [Impact Index Per Article: 2.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
Disulfide cyclization is a powerful method for reducing the conformational space of a peptide. This in turn may enable the study of its bioactive conformation. Several analogues of angiotensin II (Ang II) containing a disulfide bridge between amino acids 3 and 5 have been reported. Among these the cyclic octapeptides c[Hcy3,5]-Ang II, c[Cys3,5]-Ang II, and c[Pen 3,5]-Ang II showed significant activity at Ang II receptors. We have performed conformational analysis studies using theoretical calculations and 1H-NMR spectroscopy on tripeptide model compounds of these cyclic octapeptides which show that the cyclic moieties of c[Cys3,5]-Ang II and c[Pen3,5]-Ang II preferentially assume an inverse gamma-turn conformation. On the basis of these results, we substituted amino acid residues 3-5 in Ang II with two different gamma-turn mimetics giving four diastereomeric Ang II analogues. Interestingly, two of these are equipotent to Ang II in binding to AT1 receptors. In the contractile test using rabbit aorta rings, one of the analogues is an agonist with full contractile activity approximately equipotent to c[Pen3,5]-Ang II but 300-fold less potent than Ang II. This low potency may suggest that Ang II does not adopt a gamma-turn in the 3-5 region when interacting with the receptor.
Collapse
|
|
28 |
68 |