Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where John A. Hammer is active.

Publication


Featured researches published by John A. Hammer.


Science | 2014

Lattice light-sheet microscopy: Imaging molecules to embryos at high spatiotemporal resolution

Bi-Chang Chen; Wesley R. Legant; Kai Wang; Lin Shao; Daniel E. Milkie; Michael W. Davidson; Chris Janetopoulos; Xufeng S. Wu; John A. Hammer; Zhe Liu; Brian P. English; Yuko Mimori-Kiyosue; Daniel P. Romero; Alex T. Ritter; Jennifer Lippincott-Schwartz; Lillian K. Fritz-Laylin; R. Dyche Mullins; Diana M. Mitchell; Joshua N. Bembenek; Anne-Cécile Reymann; Ralph Böhme; Stephan W. Grill; Jennifer T. Wang; Geraldine Seydoux; U. Serdar Tulu; Daniel P. Kiehart; Eric Betzig

Introduction In vivo imaging provides a window into the spatially complex, rapidly evolving physiology of the cell that structural imaging alone cannot. However, observing this physiology directly involves inevitable tradeoffs of spatial resolution, temporal resolution, and phototoxicity. This is especially true when imaging in three dimensions, which is essential to obtain a complete picture of many dynamic subcellular processes. Although traditional in vivo imaging tools, such as widefield and confocal microscopy, and newer ones, such as light-sheet microscopy, can image in three dimensions, they sacrifice substantial spatiotemporal resolution to do so and, even then, can often be used for only very limited durations before altering the physiological state of the specimen. Lattice light-sheet microscopy. An ultrathin structured light sheet (blue-green, center) excites fluorescence (orange) in successive planes as it sweeps through a specimen (gray) to generate a 3D image. The speed, noninvasiveness, and high spatial resolution of this approach make it a promising tool for in vivo 3D imaging of fast dynamic processes in cells and embryos, as shown here in five surrounding examples. Lattice light-sheet microscopy. An ultrathin structured light sheet (blue-green, center) excites fluorescence (orange) in successive planes as it sweeps through a specimen (gray) to generate a 3D image. The speed, noninvasiveness, and high spatial resolution of this approach make it a promising tool for in vivo 3D imaging of fast dynamic processes in cells and embryos, as shown here in five surrounding examples. Rationale To address these limitations, we developed a new microscope using ultrathin light sheets derived from two-dimensional (2D) optical lattices. These are scanned plane-by-plane through the specimen to generate a 3D image. The thinness of the sheet leads to high axial resolution and negligible photobleaching and background outside of the focal plane, while its simultaneous illumination of the entire field of view permits imaging at hundreds of planes per second even at extremely low peak excitation intensities. By implementing either superresolution structured illumination or by dithering the lattice to create a uniform light sheet, we imaged cells and small embryos in three dimensions, often at subsecond intervals, for hundreds to thousands of time points at the diffraction limit and beyond. Results We demonstrated the technique on 20 different biological processes spanning four orders of magnitude in space and time, including the binding kinetics of single Sox2 transcription factor molecules, 3D superresolution photoactivated localization microscopy of nuclear lamins, dynamic organelle rearrangements and 3D tracking of microtubule plus ends during mitosis, neutrophil motility in a collagen mesh, and subcellular protein localization and dynamics during embryogenesis in Caenorhabditis elegans and Drosophila melanogaster. Throughout, we established the performance advantages of lattice light-sheet microscopy compared with previous techniques and highlighted phenomena that, when seen at increased spatiotemporal detail, may hint at previously unknown biological mechanisms. Conclusion Photobleaching and phototoxicity are typically reduced by one to two orders of magnitude relative to that seen with a 1D scanned Bessel beam or the point array scanned excitation of spinning disk confocal microscopy. This suggests that the instantaneous peak power delivered to the specimen may be an even more important metric of cell health than the total photon dose and should enable extended 3D observation of endogenous levels of even sparsely expressed proteins produced by genome editing. Improvements of similar magnitude in imaging speed and a twofold gain in axial resolution relative to confocal microscopy yield 4D spatiotemporal resolution high enough to follow fast, nanoscale dynamic processes that would otherwise be obscured by poor resolution along one or more axes of spacetime. Last, the negligible background makes lattice light-sheet microscopy a promising platform for the extension of all methods of superresolution to larger and more densely fluorescent specimens and enables the study of signaling, transport, and stochastic self-assembly in complex environments with single-molecule sensitivity. From single molecules to embryos in living color Animation defines life, and the three-dimensional (3D) imaging of dynamic biological processes occurring within living specimens is essential to understand life. However, in vivo imaging, especially in 3D, involves inevitable tradeoffs of resolution, speed, and phototoxicity. Chen et al. describe a microscope that can address these concerns. They used a class of nondiffracting beams, known as 2D optical lattices, which spread the excitation energy across the entire field of view while simultaneously eliminating out-of-focus excitation. Lattice light sheets increase the speed of image acquisition and reduce phototoxicity, which expands the range of biological problems that can be investigated. The authors illustrate the power of their approach using 20 distinct biological systems ranging from single-molecule binding kinetics to cell migration and division, immunology, and embryonic development. Science, this issue 10.1126/science.1257998 A new microscope allows three-dimensional imaging of living systems at very high resolution in real time. Although fluorescence microscopy provides a crucial window into the physiology of living specimens, many biological processes are too fragile, are too small, or occur too rapidly to see clearly with existing tools. We crafted ultrathin light sheets from two-dimensional optical lattices that allowed us to image three-dimensional (3D) dynamics for hundreds of volumes, often at subsecond intervals, at the diffraction limit and beyond. We applied this to systems spanning four orders of magnitude in space and time, including the diffusion of single transcription factor molecules in stem cell spheroids, the dynamic instability of mitotic microtubules, the immunological synapse, neutrophil motility in a 3D matrix, and embryogenesis in Caenorhabditis elegans and Drosophila melanogaster. The results provide a visceral reminder of the beauty and the complexity of living systems.


Nature Cell Biology | 2002

Identification of an organelle receptor for myosin-Va

Xufeng S. Wu; Kang Rao; Hong Zhang; Fei Wang; James R. Sellers; Lydia E. Matesic; Neal G. Copeland; Nancy A. Jenkins; John A. Hammer

Little is known about how molecular motors bind to their vesicular cargo. Here we show that myosin-Va, an actin-based vesicle motor, binds to one of its cargoes, the melanosome, by interacting with a receptor–protein complex containing Rab27a and melanophilin, a postulated Rab27a effector. Rab27a binds to the melanosome first and then recruits melanophilin, which in turn recruits myosin-Va. Melanophilin creates this link by binding to Rab27a in a GTP-dependent fashion through its amino terminus, and to myosin-Va through its carboxy terminus. Moreover, this latter interaction, similar to the ability of myosin-Va to colocalize with melanosomes and influence their distribution in vivo, is absolutely dependent on the presence of exon-F, an alternatively spliced exon in the myosin-Va tail. These results provide the first molecular description of an organelle receptor for an actin-based motor, illustrate how alternate exon usage can be used to specify cargo, and further expand the functional repertoire of Rab GTPases and their effectors.


Science | 2015

Extended-resolution structured illumination imaging of endocytic and cytoskeletal dynamics

Dong Li; Lin Shao; Bi-Chang Chen; Xi Zhang; Mingshu Zhang; Brian Moses; Daniel E. Milkie; Jordan R. Beach; John A. Hammer; Mithun Pasham; Tomas Kirchhausen; Michelle A. Baird; Michael W. Davidson; Pingyong Xu; Eric Betzig

Adding to the super-resolution arsenal Structured illumination microscopy (SIM) uses light intensities that are orders of magnitude lower than other super-resolution methods. SIM is also far faster over cellular-sized fields of view. Li et al. used two approaches to improve the resolution of SIM to allow live cell imaging of dynamic cellular processes, including endocytosis and cytoskeleton remodeling. The contrast in performance between SIM and other techniques is due to a few key differences. Defining the practical resolution at the limited signal-to-noise ratios necessary for live cell imaging will require better imaging metrics. Science, this issue 10.1126/science.aab3500 Super-resolution imaging of fast dynamic processes in living cells is facilitated by improvements to structured illumination microscopy. INTRODUCTION Various methods of super-resolution (SR) fluorescence microscopy have the potential to follow the dynamic nanoscale interactions of specific macromolecular assemblies in living cells. However, this potential is often left unfulfilled, either owing to the method’s inability to follow these processes at the speeds dictated by nature or because they require intense light that can substantially perturb the very physiology one hopes to study. An exception is structured illumination microscopy (SIM), which can image live cells far faster and with orders of magnitude less light than required for other SR approaches. However, SIM’s resolution is usually limited to only a twofold gain beyond conventional optical microscopes, or ~100 nm with visible light. RATIONALE We endeavored to find ways to extend SIM to the sub-100-nm regime while retaining, to the greatest extent possible, the advantages that make it the preferred SR method for live-cell imaging. Our first solution used an ultrahigh numerical aperture (NA) lens and total internal reflection fluorescence (TIRF) to achieve 84-nm resolution at subsecond acquisition speeds over hundreds of time points in multiple colors near the basal plasma membrane. Our second exploited the spatially patterned activation of a recently developed, reversibly photoswitchable fluorescent protein to reach 45- to 62-nm resolution, also at subsecond acquisition, over ∼10 to 40 time points. RESULTS We used high-NA TIRF-SIM to image the dynamic associations of cortical filamentous actin with myosin IIA, paxillin, or clathrin, as well as paxillin with vinculin and clathrin with transferrin receptors. Thanks to the combination of high spatial and temporal resolution, we were able to measure the sizes of individual clathrin-coated pits through their initiation, growth, and internalization. We were also able to relate pit size to lifetime, identify and characterize localized hot spots of pit generation, and describe the interaction of actin with clathrin and its role in accelerating endocytosis. With nonlinear SIM by use of patterned activation (PA NL-SIM), we monitored the remodeling of the actin cytoskeleton and the dynamics of caveolae at the cell surface. By combining TIRF-SIM and PA NL-SIM for two-color imaging, we followed the dynamic association of actin with α-actinin in expanding filopodia and membrane ruffles and characterized shape changes in and the transport of early endosomes. Last, by combining PA NL-SIM with lattice light sheet microscopy, we observed, in three dimensions and across the entire volume of whole cells, the dynamics of the actin cytoskeleton, the fusion and fission of mitochondria, and the trafficking of vesicles to and from the Golgi apparatus, each at axial resolution fivefold better than that of conventional widefield microscopy. In addition, through direct experimental comparisons, we demonstrated that the resolution for our methods is comparable with or better than other SR approaches yet allowed us to image at far higher speeds, and for far longer durations. To understand why this is so, we developed a detailed theoretical model showing that our methods transmit the information encoded in spatial frequencies beyond the diffraction limit with much greater strength than do other alternatives and hence require far fewer photons emitted from the specimen, using far less intense light. CONCLUSION High-NA TIRF-SIM and PA NL-SIM fill an unmet need for minimally invasive tools to image live cells in the gap between the 100-nm resolution traditionally associated with SIM and the sub-60-nm regime of protein-specific structural imaging served by single-molecule localization microscopy. Two approaches for improved live-cell imaging at sub-100-nm resolution. (Left) Association of cortical actin (purple) with clathrin-coated pits (green), the latter seen as rings (inset) at 84-nm resolution via a combination of total internal reflection fluorescence and structured illumination microscopy at ultrahigh numerical aperture (high-NA TIRF-SIM). (Right) Progression of resolution improvement across the actin cytoskeleton of a COS-7 cell, from conventional, diffraction-limited TIRF (220-nm resolution), to TIRF-SIM (97-nm resolution), and nonlinear SIM based on the patterned activation of a reversibly photoswitchable fluorescent protein (PA NL-SIM, 62 nm resolution). (Left and right represent single frames from time-lapse movies over 91 and 30 frames, respectively. Scale bars, 2 μm (left); 3 μm (right). Super-resolution fluorescence microscopy is distinct among nanoscale imaging tools in its ability to image protein dynamics in living cells. Structured illumination microscopy (SIM) stands out in this regard because of its high speed and low illumination intensities, but typically offers only a twofold resolution gain. We extended the resolution of live-cell SIM through two approaches: ultrahigh numerical aperture SIM at 84-nanometer lateral resolution for more than 100 multicolor frames, and nonlinear SIM with patterned activation at 45- to 62-nanometer resolution for approximately 20 to 40 frames. We applied these approaches to image dynamics near the plasma membrane of spatially resolved assemblies of clathrin and caveolin, Rab5a in early endosomes, and α-actinin, often in relationship to cortical actin. In addition, we examined mitochondria, actin, and the Golgi apparatus dynamics in three dimensions.


Nature Reviews Molecular Cell Biology | 2012

Walking to work: roles for class V myosins as cargo transporters

John A. Hammer; James R. Sellers

Cells use molecular motors, such as myosins, to move, position and segregate their organelles. Class V myosins possess biochemical and structural properties that should make them ideal actin-based cargo transporters. Indeed, studies show that class V myosins function as cargo transporters in yeast, moving a range of organelles, such as the vacuole, peroxisomes and secretory vesicles. There is also increasing evidence in vertebrate cells that class V myosins not only tether organelles to actin but also can serve as short-range, point-to-point organelle transporters, usually following long-range, microtubule-dependent organelle transport.


Current Opinion in Cell Biology | 2002

Rabs grab motors: defining the connections between Rab GTPases and motor proteins

John A. Hammer; Xufeng S. Wu

Rab GTPases and their effectors regulate membrane traffic by determining, along with cognate SNAREs, the specificity of transport vesicle docking and fusion steps. Recent studies have also implicated Rabs in the movement of these transport vesicles from their site of formation to their site of fusion, and several Rabs have been linked to specific microtubule- or actin-based motor proteins. Analyses of Rab and motor protein mutants, coupled with advanced imaging techniques, have led to the suggestion that certain Rabs function as essential components of the vesicle receptor for specific motor proteins.


Current Opinion in Cell Biology | 2010

Linking molecular motors to membrane cargo

Anna Akhmanova; John A. Hammer

Three types of motors, myosins, kinesins, and cytoplasmic dynein, cooperate to transport intracellular membrane organelles. Transport of each cargo is determined by recruitment of specific sets of motors and their regulation. Targeting of motors to membranes often depends on the formation of large multiprotein assemblies and can be influenced by membrane lipid composition. Motor activity can be regulated by cargo-induced conformational changes such as unfolding or dimerization. The architecture and function of motor: cargo complexes can also be controlled by phosphorylation, calcium signaling, and proteolysis. The complexity of transport systems is further increased by mechanical and functional cross-talk between different types of motors on the same cargo and by participation of the same motor in the movement of different organelles.


Proceedings of the National Academy of Sciences of the United States of America | 2002

Mutation of melanosome protein RAB38 in chocolate mice

Stacie K. Loftus; Denise M. Larson; Laura L. Baxter; Anthony Antonellis; Yidong Chen; Xufeng Wu; Yuan Jiang; Michael L. Bittner; John A. Hammer; William J. Pavan

Mutations of genes needed for melanocyte function can result in oculocutaneous albinism. Examination of similarities in human gene expression patterns by using microarray analysis reveals that RAB38, a small GTP binding protein, demonstrates a similar expression profile to melanocytic genes. Comparative genomic analysis localizes human RAB38 to the mouse chocolate (cht) locus. A G146T mutation occurs in the conserved GTP binding domain of RAB38 in cht mice. Rab38cht/Rab38cht mice exhibit a brown coat similar in color to mice with a mutation in tyrosinase-related protein 1 (Tyrp1), a mouse model for oculocutaneous albinism. The targeting of TYRP1 protein to the melanosome is impaired in Rab38cht/Rab38cht melanocytes. These observations, and the fact that green fluorescent protein-tagged RAB38 colocalizes with end-stage melanosomes in wild-type melanocytes, suggest that RAB38 plays a role in the sorting of TYRP1. This study demonstrates the utility of expression profile analysis to identify mammalian disease genes.


Nature | 2006

The cargo-binding domain regulates structure and activity of myosin 5

Kavitha Thirumurugan; Takeshi Sakamoto; John A. Hammer; James R. Sellers; Peter J. Knight

Myosin 5 is a two-headed motor protein that moves cargoes along actin filaments. Its tail ends in paired globular tail domains (GTDs) thought to bind cargo. At nanomolar calcium levels, actin-activated ATPase is low and the molecule is folded. Micromolar calcium concentrations activate ATPase and the molecule unfolds. Here we describe the structure of folded myosin and the GTDs role in regulating activity. Electron microscopy shows that the two heads lie either side of the tail, contacting the GTDs at a lobe of the motor domain (∼Pro 117–Pro 137) that contains conserved acidic side chains, suggesting ionic interactions between motor domain and GTD. Myosin 5 heavy meromyosin, a constitutively active fragment lacking the GTDs, is inhibited and folded by a dimeric GST–GTD fusion protein. Motility assays reveal that at nanomolar calcium levels heavy meromyosin moves robustly on actin filaments whereas few myosins bind or move. These results combine to show that with no cargo, the GTDs bind in an intramolecular manner to the motor domains, producing an inhibited and compact structure that binds weakly to actin and allows the molecule to recycle towards new cargoes.


Journal of Biological Chemistry | 2000

Effect of ADP and Ionic Strength on the Kinetic and Motile Properties of Recombinant Mouse Myosin V

Fei Wang; Lingfeng Chen; Oreste Arcucci; Estelle V. Harvey; Blair Bowers; Yuhui Xu; John A. Hammer; James R. Sellers

Mouse myosin V is a two-headed unconventional myosin with an extended neck that binds six calmodulins. Double-headed (heavy meromyosin-like) and single-headed (subfragment 1-like) fragments of mouse myosin V were expressed in Sf9 cells, and intact myosin V was purified from mouse brain. The actin-activated MgATPase of the tissue-purified myosin V, and its expressed fragments had a high V max and a lowK ATPase. Calcium regulated the MgATPase of intact myosin V but not of the fragments. Both the MgATPase activity and the in vitro motility were remarkably insensitive to ionic strength. Myosin V and its fragments translocated actin at very low myosin surface densities. ADP markedly inhibited the actin-activated MgATPase activity and the in vitromotility. ADP dissociated from myosin V subfragment 1 at a rate of about 11.5 s−1 under conditions where theV max was 3.3 s−1, indicating that, although not totally rate-limiting, ADP dissociation was close to the rate-limiting step. The high affinity for actin and the slow rate of ADP release helps the myosin head to remain attached to actin for a large fraction of each ATPase cycle and allows actin filaments to be moved by only a few myosin V molecules in vitro.


Molecular Biology of the Cell | 2012

Actin retrograde flow and actomyosin II arc contraction drive receptor cluster dynamics at the immunological synapse in Jurkat T cells

Jason Yi; Xufeng S. Wu; Travis J. Crites; John A. Hammer

Actin and myosin IIA have been implicated in the inward movement of receptor clusters at the immunological synapse of T lymphocytes. This study defines their spatial organization and quantifies their relative contributions to the dynamics of receptor clusters at the immunological synapse.

Collaboration


Dive into the John A. Hammer's collaboration.

Top Co-Authors

Avatar

Xufeng Wu

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

James R. Sellers

Boston Biomedical Research Institute

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Xufeng S. Wu

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Edward D. Korn

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Kirsten Remmert

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

Eric Betzig

Howard Hughes Medical Institute

View shared research outputs
Top Co-Authors

Avatar

Wolfgang Wagner

National Institutes of Health

View shared research outputs
Top Co-Authors

Avatar

John T. Brooks

Centers for Disease Control and Prevention

View shared research outputs
Top Co-Authors

Avatar

Jordan R. Beach

National Institutes of Health

View shared research outputs
Researchain Logo
Decentralizing Knowledge