首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
The nicotinic acetylcholine (ACh) receptor (nAChR) is the principal insecticide target. Nearly half of the insecticides by number and world market value are neonicotinoids acting as nAChR agonists or organophosphorus (OP) and methylcarbamate (MC) acetylcholinesterase (AChE) inhibitors. There was no previous evidence for in vivo interactions of the nAChR agonists and AChE inhibitors. The nitromethyleneimidazole (NMI) analog of imidacloprid, a highly potent neonicotinoid, was used here as a radioligand, uniquely allowing for direct measurements of house fly (Musca domestica) head nAChR in vivo interactions with various nicotinic agents. Nine neonicotinoids inhibited house fly brain nAChR [3H]NMI binding in vivo, corresponding to their in vitro potency and the poisoning signs or toxicity they produced in intrathoracically treated house flies. Interestingly, nine topically applied OP or MC insecticides or analogs also gave similar results relative to in vivo nAChR binding inhibition and toxicity, but now also correlating with in vivo brain AChE inhibition, indicating that ACh is the ultimate OP- or MC-induced nAChR active agent. These findings on [3H]NMI binding in house fly brain membranes validate the nAChR in vivo target for the neonicotinoids, OPs and MCs. As an exception, the remarkably potent OP neonicotinoid synergist, O-propyl O-(2-propynyl) phenylphosphonate, inhibited nAChR in vivo without the corresponding AChE inhibition, possibly via a reactive ketene metabolite reacting with a critical nucleophile in the cytochrome P450 active site and the nAChR NMI binding site.The nicotinic nervous system has two principal sites of insecticide action, the nicotinic receptor (nAChR) activated by acetylcholine (ACh) and neonicotinoid agonists (16), and acetylcholinesterase (AChE) inhibited by organophosphorus (OP) and methylcarbamate (MC) compounds to generate and maintain localized toxic ACh levels (Fig. 1) (7). The nAChR and AChE targets have been identified in insects by multiple techniques but not by direct assays of the ACh binding site in the brain of poisoned insects. Here we use the outstanding insecticidal potency of the nitromethyleneimidazole (NMI) analog of imidacloprid (IMI) (8) as a radioligand (9), designated [3H]NMI, to directly measure the house fly (Musca domestica) nAChR not only in vitro but also in vivo, allowing us to validate by a previously undescribed method the neonicotinoid direct and OP/MC indirect nAChR targets (Fig. 2). This approach also helped solve the intriguing mechanism by which an O-(2-propynyl) phosphorus compound strongly synergizes neonicotinoid insecticidal activity (10) by dual inhibition of cytochrome P450 (CYP) (1113) and the nAChR agonist site (described herein). Insecticide disruption at the insect nAChR can now be readily studied in vitro and in vivo with a single radioligand allowing better understanding of the action of several principal insecticide chemotypes (Fig. 3).Open in a separate windowFig. 1.The insect nicotinic receptor is the direct or indirect target for neonicotinoids, organophosphorus compounds and methylcarbamates, which make up about 45% of the insecticides by number and world market value (2, 7).Open in a separate windowFig. 2.In this study, Musca nicotinic receptor in vivo interactions with major insecticide chemotypes are revealed by a [3H]NMI radioligand reporter assay. *Position of tritium label.Open in a separate windowFig. 3.Two neonicotinoid nicotinic agonists and two anticholinesterase insecticides.  相似文献   

3.
4.
We used in silico methods to screen a library of 1,013 compounds for possible binding to the allosteric site in farnesyl diphosphate synthase (FPPS). Two of the 50 predicted hits had activity against either human FPPS (HsFPPS) or Trypanosoma brucei FPPS (TbFPPS), the most active being the quinone methide celastrol (IC50 versus TbFPPS ∼20 µM). Two rounds of similarity searching and activity testing then resulted in three leads that were active against HsFPPS with IC50 values in the range of ∼1–3 µM (as compared with ∼0.5 µM for the bisphosphonate inhibitor, zoledronate). The three leads were the quinone methides taxodone and taxodione and the quinone arenarone, compounds with known antibacterial and/or antitumor activity. We then obtained X-ray crystal structures of HsFPPS with taxodione+zoledronate, arenarone+zoledronate, and taxodione alone. In the zoledronate-containing structures, taxodione and arenarone bound solely to the homoallylic (isopentenyl diphosphate, IPP) site, not to the allosteric site, whereas zoledronate bound via Mg2+ to the same site as seen in other bisphosphonate-containing structures. In the taxodione-alone structure, one taxodione bound to the same site as seen in the taxodione+zoledronate structure, but the second located to a more surface-exposed site. In differential scanning calorimetry experiments, taxodione and arenarone broadened the native-to-unfolded thermal transition (Tm), quite different to the large increases in ΔTm seen with biphosphonate inhibitors. The results identify new classes of FPPS inhibitors, diterpenoids and sesquiterpenoids, that bind to the IPP site and may be of interest as anticancer and antiinfective drug leads.Farnesyl diphosphate synthase (FPPS) catalyzes the condensation of isopentenyl diphosphate (IPP; compound 1 in Fig. 1) with dimethylallyl diphosphate (DMAPP; compound 2 in Fig. 1) to form the C10 isoprenoid geranyl diphosphate (GPP; compound 3 in Fig. 1), which then condenses with a second IPP to form the C15 isoprenoid, farnesyl diphosphate (FPP; compound 4 in Fig. 1). FPP then is used in a wide range of reactions including the formation of geranylgeranyl diphosphate (GGPP) (1), squalene (involved in cholesterol and ergosterol biosynthesis), dehydrosqualene (used in formation of the Staphylococcus aureus virulence factor staphyloxanthin) (2), undecaprenyl diphosphate (used in bacterial cell wall biosynthesis), and quinone and in heme a/o biosynthesis. FPP and GGPP also are used in protein (e.g., Ras, Rho, Rac) prenylation, and FPPS is an important target for the bisphosphonate class of drugs (used to treat bone resorption diseases) such as zoledronate (compound 5 in Fig. 1) (3). Bisphosphonates targeting FPPS have activity as antiparasitics (4), act as immunomodulators (activating γδ T cells containing the Vγ2Vδ2 T-cell receptor) (5), and switch macrophages from an M2 (tumor-promoting) to an M1 (tumor-killing) phenotype (6). They also kill tumor cells (7) and inhibit angiogenesis (8). However, the bisphosphonates in clinical use (zoledronate, alendronate, risedronate, ibandronate, etidronate, and clodronate) are very hydrophilic and bind avidly to bone mineral (9). Therefore, there is interest in developing less hydrophilic species (10) that might have better activity against tumors in soft tissues and better antibacterial (11) and antiparasitic activity.Open in a separate windowFig. 1.Chemical structures of FPPS substrates, products, and inhibitors.The structure of FPPS (from chickens) was first reported by Tarshis et al. (12) and revealed a highly α-helical fold. The structures of bacterial and Homo sapiens FPPS (HsFPPS) are very similar; HsFPPS structure (13, 14) is shown in Fig. 2A. There are two substrate-binding sites, called here “S1” and “S2.” S1 is the allylic (DMAPP, GPP) binding site to which bisphosphonates such as zoledronate bind via a [Mg2+]3 cluster (15) (Fig. 2B). S2 is the homoallylic site to which IPP binds, Fig. 2B. Recently, Jahnke et al. (10) and Salcius et al. (16) discovered a third ligand-binding site called the “allosteric site” (hereafter the “A site”). A representative zoledronate+A-site inhibitor structure [Protein Data Bank (PDB) ID code 3N46] (Nov_980; compound 6 in Fig. 1) showing zoledronate in S1 and Nov_980 (compound 6) in the A site is shown in a stereo close-up view in Fig. 2B, superimposed on a zoledronate+IPP structure (PDB ID code 2F8Z) in S2. Whether the allosteric site serves a biological function (e.g., in feedback regulation) has not been reported. Nevertheless, highly potent inhibitors (IC50 ∼80 nM) have been developed (10), and the best of these newly developed inhibitors are far more hydrophobic than are typical bisphosphonates (∼2.4–3.3 for cLogP vs. ∼−3.3 for zoledronate) and are expected to have better direct antitumor effects in soft tissues (10).Open in a separate windowFig. 2.Structures of human FPPS. (A) Structure of HsFPPS showing zoledronate (compound 5) and IPP (compound 1) bound to the S1 (allylic) and S2 (homoallylic) ligand-binding sites (PDB ID code 2F8Z). (B) Superposition of the IPP-zoledronate structure (PDB ID code 2F8Z) on the zoledronate-Nov_980 A-site inhibitor structure (PDB ID code 3N46). Zoledronate binds to the allylic site S1, IPP binds to the homoallylic site S2, and the allosteric site inhibitor binds to the A site. Active-site “DDXXD” residues are indicated, as are Mg2+ molecules (green and yellow spheres, respectively). The views are in stereo.In our group we also have developed more lipophilic compounds (e.g., compound 7 in Fig. 1) (17, 18) as antiparasitic (19) and anticancer drug leads (18) and, using computational methods, have discovered other novel nonbisphosphonate FPPS inhibitors (e.g., compound 8 in Fig. 1) that have micromolar activity against FPPS (20). In this study, we extended our computational work and tried to discover other FPPS inhibitors that target the A site. Such compounds would be of interest because they might potentiate the effects of zoledronate and other bisphosphonates, as reported for other FPPS inhibitors (21), and have better tissue distribution properties in general.  相似文献   

5.
Recent archaeological research on the south coast of Peru discovered a Late Paracas (ca. 400–100 BCE) mound and geoglyph complex in the middle Chincha Valley. This complex consists of linear geoglyphs, circular rock features, ceremonial mounds, and settlements spread over a 40-km2 area. A striking feature of this culturally modified landscape is that the geoglyph lines converge on mounds and habitation sites to form discrete clusters. Likewise, these clusters contain a number of paired line segments and at least two U-shaped structures that marked the setting sun of the June solstice in antiquity. Excavations in three mounds confirm that they were built in Late Paracas times. The Chincha complex therefore predates the better-known Nasca lines to the south by several centuries and provides insight into the development and use of geoglyphs and platform mounds in Paracas society. The data presented here indicate that Paracas peoples engineered a carefully structured, ritualized landscape to demarcate areas and times for key ritual and social activities.The Chincha Valley, located 200 km south of Lima, was one of the largest and most productive regions of southern coastal Peru (Fig. 1). Previous research identified a rich prehispanic history in the valley, beginning at least in the early first millennium BCE and continuing through the Inca period in the 16th century CE (13). The earliest settled villages were part of the Paracas culture, a widespread political and social entity that began around 800 BCE and continued up to around 100 BCE. Previous field surveys identified at least 30 major Paracas period sites in the valley (1, 3, 4), making Chincha one of the main centers of development for this early Andean civilization (5). As such, it is an ideal area to test models of social evolution in general and to define the strategies that early peoples used to construct complex social organizations within the opportunities and constraints provided by their environments.Open in a separate windowFig. 1.Map showing location of the Chincha Valley, southern coastal Peru.Previous research demonstrated a dense Paracas settlement in the lower valley that focused on large platform mound complexes (Fig. 2) (4, 6). Three seasons of systematic, intensive survey and excavations by our team confirm the existence of a rich and complex Paracas occupation in the midvalley area as well, including both mound clusters and associated geoglyph features. In short, our data indicate that (i): the Chincha geoglyphs predate the better-known Nasca drainage ones by at least three centuries; (ii) Paracas period peoples created a complex landscape by constructing linear geoglyphs that converge on key settlements; and (iii) solstice marking was one component underlying the logic of geoglyph and platform mound construction and use in the Chincha Valley during the Paracas period.Open in a separate windowFig. 2.Distribution of archaeological sites linked to Paracas period settlement in Chincha, Peru. Redrawn from Canziani (4).  相似文献   

6.
Highly concise and stereospecific routes to cis and trans fusion, carrying various functionality at one of the bridgehead carbons, have been accomplished.Our group has been studying the synthetic utility of the Diels–Alder (DA) reaction of the parent cyclobutenone, 2 (1). Recently reported results demonstrate that 2 is a highly reactive, endo-selective dienophile (Fig. 1, Eq. 1) (2). We have also developed a series of intramolecular Diels–Alder (IMDA) reactions, wherein the cyclobutenone component is tethered to various conjugated dienes (compare 4); cycloaddition of these substrates delivers adducts of the type 5, which can be readily converted to trans-fused systems bearing iso-DA patterns (Fig. 1, Eq. 2, 5→6) (3). Additionally, we have described the synthesis and DA cycloaddition of an even more powerful dienophile, 2-bromocyclobutenone (Fig. 1, Eq. 3, 7) (4). The direct adducts of this [4+2] reaction (compare 8) are readily converted to norcarane carboxylic acids (9) through exposure to hydroxide base. The research described herein was initially focused on efforts to add carbon-based nucleophiles to DA cycloadducts of the type 8. It might well have been expected that such reactions would give rise to products such as 10, wherein a ketone is appended to the junction of the norcarane system (Fig. 1, Eq. 4).Open in a separate windowFig. 1.Expanding the scope of the Diels–Alder reaction.  相似文献   

7.
Repeated high-resolution bathymetric surveys of the shelf edge of the Canadian Beaufort Sea during 2- to 9-y-long survey intervals reveal rapid morphological changes. New steep-sided depressions up to 28 m in depth developed, and lateral retreat along scarp faces occurred at multiple sites. These morphological changes appeared between 120-m and 150-m water depth, near the maximum limit of the submerged glacial-age permafrost, and are attributed to permafrost thawing where ascending groundwater is concentrated along the relict permafrost boundary. The groundwater is produced by the regional thawing of the permafrost base due to the shift in the geothermal gradient as a result of the interglacial transgression of the shelf. In contrast, where groundwater discharge is reduced, sediments freeze at the ambient sea bottom temperature of ∼−1.4 °C. The consequent expansion of freezing sediment creates ice-cored topographic highs or pingos, which are particularly abundant adjacent to the discharge area.

The effects of on-going terrestrial permafrost degradation (13) have been appraised by comparison of sequential images of Arctic landscapes that show geomorphic changes attributed primarily to thermokarst activity induced by recent atmospheric warming and ongoing natural periglacial processes (48). While the existence of extensive relict submarine permafrost on the continental shelves in the Arctic has been known for years (9, 10), the dynamics of submarine permafrost growth and decay and consequent modifications of seafloor morphology are largely unexplored.Throughout the Pleistocene, much of the vast continental shelf areas of the Arctic Ocean experienced marine transgressions and regressions associated with ∼125-m global sea level changes (11). Extensive terrestrial permafrost formed during sea-level low stands when the mean annual air temperatures of the exposed shelves were less than −15 °C (11, 12). Exploration wells drilled on the continental shelf in the Canadian Beaufort Sea show that relict terrestrial permafrost occurs in places to depths >600 m below seafloor (mbsf) and forms a seaward-thinning wedge beneath the outer shelf (10, 13, 14) (Fig. 1 A and B). The hydrography of the Canadian Beaufort Sea slows the degradation of the relict permafrost because a cold-water layer with temperatures usually near -1.4 °C blankets the seafloor from midshelf depths down to ∼200-m water depth (mwd) (15, 16) (Fig. 1B). As the freezing point temperature of interstitial waters is also controlled by salinity and sediment grain size, partially frozen sediments occur in a zone delimited by the ∼−2 °C and 0 °C isotherms (17, 18) (Fig. 1B).Open in a separate windowFig. 1.Map and cross-section showing the relationship between shelf edge morphology and the subsurface thermal structure along the shelf edge in the Canadian Beaufort Sea. (A) Shows the location of the study area with respect to estimates of submarine permafrost density (see key) and thickness, modified after 14. Thin contours indicate permafrost thickness in meters. Thicker contour is 120-m isobath marking the shelf edge. Area of repeat mapping coverage shown in Fig. 2 is indicated with a red box. (B) Shows a schematic cross-section with contours of selected subsurface isotherms modified after 15 along line x-x’ in A. The dotted blue line illustrates a thermal minimum (T-min) running through the relict permafrost isotherm and Beaufort Sea waters (16). Green shading indicates relict permafrost. Turquoise arrows show inferred flow of water from permafrost thawing along the base of the relict permafrost to the seafloor. The brown area indicates the zone where relict Pleistocene permafrost is predicted to have thawed with consequent movement of liberated groundwater, associated latent heat transfer and thaw consolidation causing surface settlement. Dashed brown lines define the subbottom limits for methane hydrate stability zone (MHSZ) which starts at ∼240 m below the sea surface and extends into the subsurface depending on the pressure and temperature gradient. The red box indicates the area shown in more detail in C with the same color scheme. The area of denuded seafloor in C is flanked by PLFs (dark-blue fill). Red arrows indicate the direction of heat transfer along the seaward edge of relict permafrost wedge.Distinctive surface morphologies characterize terrestrial permafrost areas. Conical hills (3 to 100 m in diameter) called pingos are common in the Arctic (19, 20). Pingos are formed due to freezing of groundwater. They characteristically contain lenses of nearly pure ground ice that cause heaving of the ground surface. Positive relief features with similar dimensions, referred to as pingo-like features (PLFs), are scattered across the Canadian Beaufort shelf (21, 22). On land, permafrost thawing, where there is ground ice in excess of the sediment pore space, can induce sediment consolidation (23), and surface subsidence results in widespread thermokarst landforms. Among the more dramatic occurrences are retrogressive thaw slumps (48, 24). These form where ice-rich permafrost experiences surface thaw causing thaw settlement and release of liquified sediment flows. Because of the loss of volume associated with thawing of massive ground ice, thaw slumps can quickly denude permafrost landscapes.During the first systematic multibeam mapping surveys in 2010 covering part of the shelf edge and slope in the Canadian Beaufort Sea, a band of unusually rough seafloor morphology between ∼120 and ∼200 mwds (25) was discovered along a ∼95-km-long stretch of the shelf. Subsequently, three additional multibeam surveys covering small characteristic areas (Fig. 2A) were conducted to understand the processes responsible for the observed morphologies. Here, we document the unique morphologies and seafloor change in this area and explore how the seafloor features may be related to subsea permafrost degradation and formation.Open in a separate windowFig. 2.(A) Shows bathymetry of a small section of the shelf edge indicated in Fig. 1A, with a color scale going from white (128 m) to blue (200 m) and contours at 120, 140, 170, and 200 mbsf. Outlines of areas resurveyed in 2013 (blue), 2017 (turquoise), and 2019 (green) are superimposed on the 2019 and regional 2010 survey. Colored symbols indicate locations of cores with porewater data using same key as Fig. 5B. The red star indicates the location of a temperature tripod deployed in the period 2015 to 2016. The location of ROV dive tracks (blue paths) are indicated. (B) Covers the same area as A with polygons identifying sites where changes were noted between surveys as follows: 2010 to 2013 (green), 2013 to 2017 (purple), 2017 to 2019 (black), and 2010 and 2019 (red). (C) Shows the same area, colored according to the difference in bathymetry between the 2019 survey and an idealized smooth surface extending between the top of the shelf edge scarp and the layered sediments occurring between the numerous PLFs. This is used to estimate the volume of material that eroded assuming the earlier Holocene seafloor corresponded with this idealized surface. Three zones of topography are labeled. Red boxes are locations of Figs. 3 and and5.5. (D) Shows Chirp profiles with the position of profiles shown in C and Fig. 5A. Light-green backdrop in X-X’ indicates possible void produced by retrogressive slide retreat used to calculated volume loss. Also indicated are TL, tilted layers; P, pingo-like-feature; and DR, diffuse reflector.  相似文献   

8.
Recently, a type of ferroelectric nematic fluid has been discovered in liquid crystals in which the molecular polar nature at molecule level is amplified to macroscopic scales through a ferroelectric packing of rod-shaped molecules. Here, we report on the experimental proof of a polar chiral liquid matter state, dubbed helielectric nematic, stabilized by the local polar ordering coupled to the chiral helicity. This helielectric structure carries the polar vector rotating helically, analogous to the magnetic counterpart of helimagnet. The helielectric state can be retained down to room temperature and demonstrates gigantic dielectric and nonlinear optical responses. This matter state opens a new chapter for developing the diverse polar liquid crystal devices.

In nature , a new matter state usually arises as a result of unexpected combinations of hierarchical orderings. Helicity is one of the most essential nature of matter states for organizing superstructures in soft matters, spanning many length scales from the atomic to the macroscopic biological levels. When constructed from building blocks with inherent polarity, three hierarchical orderings could coexist in a helical structure: 1) the head-to-tail or polar symmetry of each building block (e.g., Fig. 1C), 2) the orientational order of a swarm of building blocks (Fig. 1A), and 3) the emergent helicity (Fig. 1B). While a simultaneous realization of these three orderings could lead to extraordinary material properties, such highly hierarchical structures are often challenging to achieve in man-made systems. Probably the most familiar example is the chiral magnet or helimagnet (Fig. 1B) in quantum systems, where the magnetic spins form two- or three-dimensional spiral structures (1, 2). The polar magnetic helical structures are considered mainly to originate from either the breaking of the space-inversion symmetry in crystal structures (3) or the magnetic frustration (1, 4, 5). Their strong magnetism-chirality coupling triggers enormous interests in condensed matter physics, leading to many unique quantum and information functionalities (69). From the mirror relationship between the magnetism and electricity, we anticipate the incidence of a possible electric version of the helimagnets, namely helielectrics. However, the diverse magnetic topological states rarely show up in electric systems, except a few recent breakthroughs (e.g., the observation of the electric skyrmions, polar vortices, and merons in metal-organic crystals) (1012). The special electric states at nanoscale exhibit extraordinary properties such as local negative dielectric permittivity (13) and strain-polarization coupling (14, 15). Nevertheless, nearly all the aforementioned chiral magnet or electric-analog systems are based on elaborately fabricated inorganics. It is expected that the revolutionary realization of these topologies in a soft matter system would bring the advantages of flexibility, simple preparation, large-area film formation, and ease of integration into electric devices.Open in a separate windowFig. 1.Topological analogy: electric versus magnetic states. (A) Uniform magnetization or polarization. (B) Helimagnet or helielectric states. Possible helicoidal (top) and heliconical (bottom) textures are shown. (C) Molecular structure of the polar anisotropic entity, RM734. The molecular polar dipole is nearly parallel to the long molecular axis. (D) The ferroelectric nematic state with spontaneous polarization. (E) HN* state with heli realized by adding chiral generators into the polar chiral nematic state. One-dimensional polarization fields are also depicted in D and E for clarity. (F) The molecular structures of the chiral generators S1 and S2. (G) The state diagram of the two HN* materials by mixing RM734 with S1 or S2.Among the soft matter systems, liquid analogs of ferromagnet and helimagnet have been reported in liquid crystal (LC) colloids recently (1620). For the electric versions, there already exist a category of materials possessing all the aforementioned three hierarchical orderings (i.e., the ferroelectric smectic LCs) (2126). The smectic C* (SmC*) has layered heliconical structure with its local polarity aligning perpendicular to the long molecular axis. Confinement to thin LC cells leads to the unwound ferroelectric state of SmC* with microsecond switching time, thereby being a promising candidate for LC display applications. However, the unavoidable defect generation in the devices originated from the crystal-like structure has been one of the main technical difficulties. Moreover, the SmC* has intrinsically low fluidity and polarity (spontaneous polarization Ps < 1 μC). Here, we report a discovery of a helimagnetic analog state in polar LC materials, dubbed helielectric nematic (HN*). The spontaneous polar nematic ordering is coupled to the chiral orientational helicity (Fig. 1B), taking the form with a nearly helicoidal orientational field. Thanks to its much higher fluidity than the traditional SmC* ferroelectrics, uniform structures can be easily obtained by the typical thermal annealing process. The simultaneous observation of the traditional nonlinear second-harmonic generation (SHG) and SHG interferometry microscopies, as well as the optical observations of the selective reflection from HN* state, allow us to directly visualize the helical polar field. In contrast to the traditional nanoscopic helimagnetic or helielectric inorganics, a wide tunability of the periodic distance ranging from micrometers to near ultraviolet wavelength is achieved in the fluidic structure. Besides, the ability of switching between the polar and nonpolar helical LC states enables complementary physics study for the topology features in HN*. As gifts of the chirality–polarity interaction, the matter state uniquely expresses giant dielectric and SHG optical response, especially interesting SHG amplification when the SHG wavelength coincides with the reflection band of the HN* state.  相似文献   

9.
A three-dimensionally preserved 2-mm-long larva of the arthropod Leanchoilia illecebrosa from the 520-million-year-old early Cambrian Chengjiang biota of China represents the first evidence, to our knowledge, of such an early developmental stage in a short-great-appendage (SGA) arthropod. The larva possesses a pair of three-fingered great appendages, a hypostome, and four pairs of well-developed biramous appendages. More posteriorly, a series of rudimentary limb Anlagen revealed by X-ray microcomputed tomography shows a gradient of decreasing differentiation toward the rear. This, and postembryonic segment addition at the putative growth zone, are features of late-stage metanauplii of eucrustaceans. L. illecebrosa and other SGA arthropods, however, are considered representative of early chelicerates or part of the stem lineage of all euarthropods. The larva of an early Cambrian SGA arthropod with a small number of anterior segments and their respective appendages suggests that posthatching segment addition occurred in the ancestor of Euarthropoda.Evolutionary developmental biology (evo-devo) explains evolutionary changes in different organisms by investigating their developmental processes (1). Paleontology contributes to evo-devo by providing information that is only available in fossil organisms (2). Studies of evolutionary development in fossil arthropods, which have dominated faunas from the early Cambrian (∼520 million years ago) to the present, have focused on trilobites (3), “Orsten”-type fossil crustaceans (46), and Mesozoic malacostracan crustaceans (7). Due to their small size and low preservation potential, fossil evidence of the appendages of early developmental stages of arthropods are rare, and known mainly from those with the special “Orsten” type of preservation (8), i.e., with the cuticle secondarily phosphatized, from the mid-Cambrian (500–497 million years ago) (9).Here we describe an exceptionally preserved early developmental stage of a Cambrian arthropod from the Chengjiang biota of China. The specimen is only 2 mm long and is three-dimensionally preserved (Fig. 1, Insets). We interpret this specimen as a representative of the short-great-appendage (SGA) arthropod Leanchoilia illecebrosa—the most abundant SGA arthropod from this biota (10). SGA arthropods form a distinct early group characterized by prominent anteriormost appendages specialized for sensory (11) or feeding purposes (11, 12). Thus far, knowledge of L. illecebrosa is based mainly on adult specimens with a body length ranging from 20 to 46 mm (13) (Fig. 1). Specimens smaller than 20 mm are rare—only two examples, both 8 mm long, have been reported (8, 12) (Fig. S1B).Open in a separate windowFig. 1.L. illecebrosa from the Chengjiang biota. Macrophotographs of an adult (specimen YKLP 11087) and the minute larva (Insets; specimen YKLP 11088a, b). cs, cephalic shield; rs, rostrum; sga, short great appendage; ts1 and ts11, trunk segments 1 and 11; te, telson. Insets are to the same scale as main image. (Scale bar: 5 mm.)Open in a separate windowFig. S1.Two larval stages of L. illecebrosa. (A) The 2-mm-long larva described here (specimen YKLP 11088a, b). (B) An 8-mm-long larva previously reported in ref. 12 (specimen YKLP 11084a, b; reprinted with permission from ref. 12). (Scale bar: 2 mm.)  相似文献   

10.
In the middle-to-late Holocene, Earth’s monsoonal regions experienced catastrophic precipitation decreases that produced green to desert state shifts. Resulting hydrologic regime change negatively impacted water availability and Neolithic cultures. Whereas mid-Holocene drying is commonly attributed to slow insolation reduction and subsequent nonlinear vegetation–atmosphere feedbacks that produce threshold conditions, evidence of trigger events initiating state switching has remained elusive. Here we document a threshold event ca. 4,200 years ago in the Hunshandake Sandy Lands of Inner Mongolia, northern China, associated with groundwater capture by the Xilamulun River. This process initiated a sudden and irreversible region-wide hydrologic event that exacerbated the desertification of the Hunshandake, resulting in post-Humid Period mass migration of northern China’s Neolithic cultures. The Hunshandake remains arid and is unlikely, even with massive rehabilitation efforts, to revert back to green conditions.Earth’s climate is subject to abrupt, severe, and widespread change, with nonlinear vegetation–atmosphere feedbacks that produced extensive and catastrophic ecosystem shifts and subsequent cultural disruption and dispersion during the Holocene (17). In the early and middle Holocene, northern China’s eastern deserts, including much of the currently sparsely vegetated and semistabilized Hunshandake (Figs. 1 and and2),2), were covered by forests (8), reflecting significantly wetter climate associated with intensification of monsoon precipitation by up to 50% (6).Open in a separate windowFig. 1.Geographical location of the Hunshandake Sandy Lands (A) and its area (encircled by red line in B). The black rectangle in B marks the location of the enlarged maps C and D on the Right, and the green rectangle shows the location of Fig. 2. Map C shows the localities of water samples, and map D shows the localities of sections with stratigraphy presented in Fig. 3. The sand–paleosol section P (Fig. 3) is on the southern margin, and the site Bayanchagan marks the coring site of ref. 8. Rivers with headwaters in the Hunshandake likely formed by groundwater sapping are marked in blue. Drainages to the southwest and west are currently undergoing groundwater sapping, with substantial spring-driven flow found at the current river base level.Open in a separate windowFig. 2.(Left) Holocene lakes and channels in the Hunshandake and lake extent at selected epochs. Upper, middle, and lower lakes are indicated by points A, B, and C, respectively. Xilamulun River (point D) drains to the east. Groundwater-sapping headcuts at the upper reaches of incised canyons (point E) suggest a mid-Holocene interval of easterly surface flow, followed by groundwater drainage beginning at the ca. 4.2 ka event. Northern and central channels at point E are currently abandoned, and groundwater sapping has migrated to the southerly of the three channels shown. (Right) Cross-sections of the predrainage shift, northerly drainage into Dali Lake (Localities shown on the Left), showing the increase in widths of channels downstream (Vertical exaggeration ∼30:1).Monsoonal weakening, in response to middle-to-late Holocene insolation decrease, reduced precipitation, leading to a green/sandy shift and desertification across Inner Mongolia between ca. 5,000 and 3,000 y (years) ago (6). However, variations in the timing of this transition (9, 10) suggest local/regional thresholds or possibly environmental tipping by stochastic fluctuations. The impacts of this wet-to-dry shift in the Hunshandake, expressed as variations in surface and subsurface hydrology coincident with the termination of the formation of thick and spatially extensive paleosols, and the impacts of a ca. 4.2 ka (1 ka = 1,000 years) mid-Holocene desiccation of the Hunshandake on the development of early Chinese culture remain poorly understood and controversial (6, 11). Here we report for the first time to our knowledge on variability in a large early-to-middle Holocene freshwater lake system in China’s Hunshandake Sandy Lands and associated vegetation change, which demonstrates a model of abrupt green/desert switching. We document a possible hydrologic trigger event for this switching and discuss associated vegetation and hydrologic disruptions that significantly impacted human activities in the region.  相似文献   

11.
Since Darwin, biologists have been struck by the extraordinary diversity of teleost fishes, particularly in contrast to their closest “living fossil” holostean relatives. Hypothesized drivers of teleost success include innovations in jaw mechanics, reproductive biology and, particularly at present, genomic architecture, yet all scenarios presuppose enhanced phenotypic diversification in teleosts. We test this key assumption by quantifying evolutionary rate and capacity for innovation in size and shape for the first 160 million y (Permian–Early Cretaceous) of evolution in neopterygian fishes (the more extensive clade containing teleosts and holosteans). We find that early teleosts do not show enhanced phenotypic evolution relative to holosteans. Instead, holostean rates and innovation often match or can even exceed those of stem-, crown-, and total-group teleosts, belying the living fossil reputation of their extant representatives. In addition, we find some evidence for heterogeneity within the teleost lineage. Although stem teleosts excel at discovering new body shapes, early crown-group taxa commonly display higher rates of shape evolution. However, the latter reflects low rates of shape evolution in stem teleosts relative to all other neopterygian taxa, rather than an exceptional feature of early crown teleosts. These results complement those emerging from studies of both extant teleosts as a whole and their sublineages, which generally fail to detect an association between genome duplication and significant shifts in rates of lineage diversification.Numbering ∼29,000 species, teleost fishes account for half of modern vertebrate richness. In contrast, their holostean sister group, consisting of gars and the bowfin, represents a mere eight species restricted to the freshwaters of eastern North America (1). This stark contrast between teleosts and Darwin''s original “living fossils” (2) provides the basis for assertions of teleost evolutionary superiority that are central to textbook scenarios (3, 4). Classic explanations for teleost success include key innovations in feeding (3, 5) (e.g., protrusible jaws and pharyngeal jaws) and reproduction (6, 7). More recent work implicates the duplicate genomes of teleosts (810) as the driver of their prolific phenotypic diversification (8, 1113), concordant with the more general hypothesis that increased morphological complexity and innovation is an expected consequence of genome duplication (14, 15).Most arguments for enhanced phenotypic evolution in teleosts have been asserted rather than demonstrated (8, 11, 12, 15, 16; but see ref. 17), and draw heavily on the snapshot of taxonomic and phenotypic imbalance apparent between living holosteans and teleosts. The fossil record challenges this neontological narrative by revealing the remarkable taxonomic richness and morphological diversity of extinct holosteans (Fig. 1) (18, 19) and highlights geological intervals when holostean taxonomic richness exceeded that of teleosts (20). This paleontological view has an extensive pedigree. Darwin (2) invoked a long interval of cryptic teleost evolution preceding the late Mesozoic diversification of the modern radiation, a view subsequently supported by the implicit (18) or explicit (19) association of Triassic–Jurassic species previously recognized as “holostean ganoids” with the base of teleost phylogeny. This perspective became enshrined in mid-20th century treatments of actinopterygian evolution, which recognized an early-mid Mesozoic phase dominated by holosteans sensu lato and a later interval, extending to the modern day, dominated by teleosts (4, 20, 21). Contemporary paleontological accounts echo the classic interpretation of modest teleost origins (2224), despite a systematic framework that substantially revises the classifications upon which older scenarios were based (2225). Identification of explosive lineage diversification in nested teleost subclades like otophysans and percomorphs, rather than across the group as a whole, provides some circumstantial neontological support for this narrative (26).Open in a separate windowFig. 1.Phenotypic variation in early crown neopterygians. (A) Total-group holosteans. (B) Stem-group teleosts. (C) Crown-group teleosts. Taxa illustrated to scale.In contrast to quantified taxonomic patterns (20, 23, 24, 27), phenotypic evolution in early neopterygians has only been discussed in qualitative terms. The implicit paleontological model of morphological conservatism among early teleosts contrasts with the observation that clades aligned with the teleost stem lineage include some of the most divergent early neopterygians in terms of both size and shape (Fig. 1) (see, for example, refs. 28 and 29). These discrepancies point to considerable ambiguity in initial patterns of phenotypic diversification that lead to a striking contrast in the vertebrate tree of life, and underpins one of the most successful radiations of backboned animals.Here we tackle this uncertainty by quantifying rates of phenotypic evolution and capacity for evolutionary innovation for the first 160 million y of the crown neopterygian radiation. This late Permian (Wuchiapingian, ca. 260 Ma) to Cretaceous (Albian, ca. 100 Ma) sampling interval permits incorporation of diverse fossil holosteans and stem teleosts alongside early diverging crown teleost taxa (Figs. 1 and and2A2A and Figs. S1 and andS2),S2), resulting in a dataset of 483 nominal species-level lineages roughly divided between the holostean and teleost total groups (Fig. 2B and Fig. S2). Although genera are widely used as the currency in paleobiological studies of fossil fishes (30; but see ref. 31), we sampled at the species level to circumvent problems associated with representing geological age and morphology for multiple congeneric lineages. We gathered size [both log-transformed standard length (SL) and centroid size (CS); results from both are highly comparable (Figs. S3 and andS4);S4); SL results are reported in the main text] and shape data (the first three morphospace axes arising from a geometric morphometric analysis) (Fig. 2A and Figs. S1) from species where possible. To place these data within a phylogenetic context, we assembled a supertree based on published hypotheses of relationships. We assigned branch durations to a collection of trees under two scenarios for the timescale of neopterygian diversification based on molecular clock and paleontological estimates. Together, these scenarios bracket a range of plausible evolutionary timelines for this radiation (Fig. 2B). We used the samples of trees in conjunction with our morphological datasets to test for contrasts in rates of, and capacity for, phenotypic change between different partitions of the neopterygian Tree of Life (crown-, total-, and stem-group teleosts, total-group holosteans, and neopterygians minus crown-group teleosts), and the sensitivity of these conclusions to uncertainty in both relationships and evolutionary timescale. Critically, these include comparisons of phenotypic evolution in early crown-group teleosts—those species that are known with certainty to possess duplicate genomes—with rates in taxa characterized largely (neopterygians minus crown teleosts) or exclusively (holosteans) by unduplicated genomes. By restricting our scope to early diverging crown teleost lineages, we avoid potentially confounding signals from highly nested radiations that substantially postdate both genome duplication and the origin of crown teleosts (26, 32). This approach provides a test of widely held assumptions about the nature of morphological evolution in teleosts and their holostean sister lineage.Open in a separate windowFig. 2.(A) Morphospace of Permian–Early Cretaceous crown Neopterygii. (B) One supertree subjected to our paleontological (Upper) and molecular (Lower) timescaling procedures to illustrate contrasts in the range of evolutionary timescales considered. Colors of points (A) and branches (B) indicate membership in major partitions of neopterygian phylogeny. Topologies are given in Datasets S4 and S5. See Dataset S6 for source trees.Open in a separate windowFig. S1.Morphospace of 398 Permian–Early Cretaceous Neopterygii. Three major axes of shape variation are presented. Silhouettes and accompanying arrows illustrate the main anatomical correlates of these principal axes, as described in Open in a separate windowFig. S2.Morphospace of 398 Permian–Early Cretaceous Neopterygii, illustrating the major clades of (A) teleosts and (B) holosteans.Open in a separate windowFig. S3.Comparisons of size rates between (A) holosteans and teleosts, (B) crown teleosts and all other neopterygians, (C) crown teleosts and stem teleosts, (D) crown teleosts and holosteans, and (E) stem teleosts and holosteans. Comparisons were made using the full-size SL dataset, a CS dataset, and a smaller SL dataset pruned to exactly match the taxon sampling of the CS dataset. Identical taxon sampling leads the CS and pruned SL datasets to yield near identical results. Although the larger SL dataset results often differ slightly, the overall conclusion from each pairwise comparison (i.e., which outcome is the most likely in an overall majority of trees) is identical in all but one comparison (E, under molecular timescales).Open in a separate windowFig. S4.Comparisons of size innovation between (A) holosteans and teleosts, (B) crown teleosts and all other neopterygians, (C) crown teleosts and stem teleosts, (D) crown teleosts and holosteans, and (E) stem teleosts and holosteans. Comparisons were made using the full-size SL dataset, a CS dataset, and a smaller SL dataset pruned to exactly match the taxon sampling of the CS dataset. Comparisons of size innovation are presented for K value distributions of the three datasets resemble each other closely.  相似文献   

12.
13.
The genome of transposable phage Mu is packaged as a linear segment, flanked by several hundred base pairs of non-Mu DNA. The linear ends are held together and protected from nucleases by the phage N protein. After transposition into the Escherichia coli chromosome, the flanking DNA (FD) is degraded, and the 5-bp gaps left in the target are repaired to generate a simple Mu insertion. Our study provides insights into this repair pathway. The data suggest that the first event in repair is removal of the FD by the RecBCD exonuclease, whose entry past the N-protein block is licensed by the transpososome. In vitro experiments reveal that, when RecBCD is allowed entry into the FD, it degrades this DNA until it arrives at the transpososome, which presents a barrier for further RecBCD movement. RecBCD action is required for stimulating endonucleolytic cleavage within the transpososome-protected DNA, leaving 4-nt flanks outside both Mu ends. This end product of collaboration between the transpososome and RecBCD resembles the intermediate products of Tn7 and retroviral and retrotransposon transposition, and may hint at a common gap-repair mechanism in these diverse transposons.The repair of transposon insertions is the least understood aspect of transposon biology. Transposable phage Mu provides an excellent system to study this process because of its high transposition frequency. Mu is a temperate phage, which uses transposition not only to integrate into its Escherichia coli host to generate prophages but also to amplify its genome during the lytic cycle, where transposition is coupled to replication (Fig. 1A) (14).Open in a separate windowFig. 1.Two transposition pathways during the Mu life cycle. The chemical steps of single-stranded DNA cleavage at Mu ends followed by strand transfer (ST) of the cleaved ends to phosphodiester bonds spaced 5 bp apart on the target are the same in both the lytic (A) and infection (B) phases of transposition. During the lytic phase, Mu transposition is intramolecular, and the ST intermediate is resolved by target-primed replication through Mu (A). During the infection phase, transposition is intermolecular, and the Ɵ ST intermediate is resolved by removal of the flanking DNA and repair of the 5-bp gaps left in the target (B). The FD is covalently closed in A but noncovalently closed by phage N protein (oval) bound to the tips of the FD in B. The target DNA flanking Mu is red in all of the figures. Arrowheads indicate the 3′ ends.This study focuses on the transposition event that occurs immediately after infection, which is followed by repair rather than replication of the Mu insertion. The infecting Mu genome is linear, and is linked to several hundred base pairs of non–Mu-flanking DNA. The tips of this DNA are held together by an injected phage protein, N, which converts it into a noncovalently closed circular form and protects the DNA against exonucleases (57). After integration, the flanking DNA (FD) is removed (8), and the 5-bp gaps generated in the target are repaired to yield a simple Mu insertion (Fig. 1B) (9). Unlike the noncovalently closed configuration of the infecting Mu genome, replicating Mu is initially part of a covalently closed circular E. coli chromosome (Fig. 1A). At the end of the lytic phase, Mu replicas are packaged such that host DNA linked to either side of the insertion is included in the phage head; this is the source of the FD in the infecting Mu genome.The chemical steps of Mu transposition during the replication and repair pathways are the same, namely the MuA transposase nicks Mu DNA at each end and then joins the nicked ends to target DNA cleaved 5 bp apart (Fig. 1) (2, 8). However, the resulting branched strand-transfer intermediate is resolved alternatively by target-primed replication during the lytic phase (Fig. 1A) (10) and by FD removal and limited replicative repair during the infection phase (Fig. 1B) (8, 11). The alternate fates of a similar strand-transfer joint have long been a matter of speculation. Although much has been learned about the proteins that promote the transition from transposition to replication (10, 12, 13), little is known about those that assist in repair, other than that the gap-filling E. coli polymerase PolA is not required and that the double-strand break (DSB) repair machinery is somehow involved (14).We show in this study that the noncovalently closed configuration of the FD in the infecting Mu donor substrate controls the fate of the strand-transfer joint. Our data show that the RecBCD exonuclease is required in vivo for degradation of the FD. This occurs only after integration of infecting Mu, suggesting that there is a timed mechanism to remove N and allow RecBCD access into the FD. Attempts to recapitulate this reaction in vitro have revealed that the FD is processed in two stages: First, RecBCD degrades the long DNA, and next, a specific endonucleolytic cleavage leaves short 4-nt flanks within the transpososome-protected DNA. Earlier, we had found that a cryptic endonuclease activity of MuA was required for removal of the flanks in vivo, and had interpreted the data to suggest that degradation of this DNA is initiated by endonucleolytic cleavage (15). In light of the results obtained in the present study, we reinterpret our earlier data and propose a new model for repair. We expect that this first biochemical analysis, to our knowledge, of the initial steps of Mu repair will reveal pathways common to the repair of all transposon insertions.  相似文献   

14.
15.
Aryl chlorides are among the most versatile synthetic precursors, and yet inexpensive and benign chlorination techniques to produce them are underdeveloped. We propose a process to generate aryl chlorides by chloro-group transfer from chlorophenol pollutants to arenes during their mineralization, catalyzed by Cu(NO3)2/NaNO3 under aerobic conditions. A wide range of arene substrates have been chlorinated using this process. Mechanistic studies show that the Cu catalyst acts in cooperation with NOx species generated from the decomposition of NaNO3 to regulate the formation of chlorine radicals that mediate the chlorination of arenes together with the mineralization of chlorophenol. The selective formation of aryl chlorides with the concomitant degradation of toxic chlorophenol pollutants represents a new approach in environmental pollutant detoxication. A reduction in the use of traditional chlorination reagents provides another (indirect) benefit of this procedure.

Chlorophenols are widely encountered moieties present in herbicides, drugs, and pesticides (1). Owing to the high dissociation energies of carbon‒chloride bonds, chlorophenols biodegrade very slowly, and their prolonged exposure leads to severe ecological and environmental problems (Fig. 1A) (24). Several well-established technologies have been developed for the treating of chlorophenols, including catalytic oxidation (511), biodegradation (1215), solvent extraction (16, 17), and adsorption (1820) Among these methods, adsorption is the most versatile and widely used method due to its high removal efficiency and simple operation, but the resulting products are of no value, and consequently, these processes are not viable.Open in a separate windowFig. 1.Background and reaction design. (A) Examples of chlorophenol pollutants. (B) Examples of aryl chlorides. (C) The chlorination process reported herein was based on chloro-group transfer from chlorophenol pollutants.With the extensive application of substitution reactions (21, 22), transfunctionalizations (23, 24), and cross-coupling reactions (25, 26), aryl chlorides are regarded as one of the most important building blocks widely used in the manufacture of polymers, pharmaceuticals, and other types of chemicals and materials (Fig. 1B) (2731). Chlorination of arenes is usually carried out with toxic and corrosive reagents (3234). Less toxic and more selective chlorination reagents tend to be expensive [e.g., chloroamides (35, 36)] and are not atom economic (3739). Consequently, from the perspective of sustainability, the ability to transfer a chloro group from unwanted chlorophenols to other substrates would be advantageous.Catalytic isofunctional reactions, including transfer hydrogenation and alkene metathesis, have been widely exploited in organic synthesis. We hypothesized that chlorination of arenes also could be achieved by chloro-group transfer, and since stockpiles of chlorophenols tend to be destroyed by mineralization and chlorophenol pollutants may be concentrated by adsorption (1820), they could be valorized as chlorination reagents via transfer of the chloro group to arene substrates during their mineralization, thereby adding value to the destruction process (Fig. 1C). Although chlorophenol pollutants are not benign, their application as chlorination reagents, with their concomitant destruction to harmless compounds, may be considered as not only meeting the criteria of green chemistry but also potentially surpassing it. Herein, we describe a robust strategy to realize chloro-group transfer from chlorophenol pollutants to arenes and afford a wide range of value-added aryl chlorides.  相似文献   

16.
17.
The availability of plants and freshwater shapes the diets and social behavior of chimpanzees, our closest living relative. However, limited evidence about the spatial relationships shared between ancestral human (hominin) remains, edible resources, refuge, and freshwater leaves the influence of local resources on our species’ evolution open to debate. Exceptionally well-preserved organic geochemical fossils—biomarkers—preserved in a soil horizon resolve different plant communities at meter scales across a contiguous 25,000 m2 archaeological land surface at Olduvai Gorge from about 2 Ma. Biomarkers reveal hominins had access to aquatic plants and protective woods in a patchwork landscape, which included a spring-fed wetland near a woodland that both were surrounded by open grassland. Numerous cut-marked animal bones are located within the wooded area, and within meters of wetland vegetation delineated by biomarkers for ferns and sedges. Taken together, plant biomarkers, clustered bone debris, and hominin remains define a clear spatial pattern that places animal butchery amid the refuge of an isolated forest patch and near freshwater with diverse edible resources.Spatial patterns in archaeological remains provide a glimpse into the lives of our ancestors (15). Although many early hominin environments are interpreted as grassy or open woodlands (68), fossil bones and plant remains are rarely preserved together in the same settings. As a result, associated landscape reconstructions commonly lack coexisting fossil evidence for hominins and local-scale habitat (microhabitat) that defined the distribution of plant foods, refuge, and water (7). This problem is exacerbated by the discontinuous nature and low time resolution often available across ancient soil (paleosol) horizons, including hominin archaeological localities. One notable exception is well-time-correlated 1.8-million-y-old paleosol horizons exposed at Olduvai Gorge. Associated horizons contain exceptionally preserved plant biomarkers along with many artifacts and fossilized bones. Plant biomarkers, which previously revealed temporal patterns in vegetation and water (8), are well preserved in the paleosol horizon and document plant-type spatial distributions that provide an ecosystem context (9, 10) for resources that likely affected the diets and behavior of hominin inhabitants.Plant biomarkers are delivered by litter to soils and can distinguish plant functional type differences in standing biomass over scales of 1–1,000 m2 (11). Trees, grasses, and other terrestrial plants produce leaf waxes that include long-chain n-alkanes such as hentriacontane (nC31), whereas aquatic plants and phytoplankton produce midchain homologs (e.g., nC23) (12, 13). The ratio of shorter- versus long-chain n-alkane abundances distinguish relative organic matter inputs from aquatic versus terrestrial plants to sediments (13):Paq = (nC23nC25)/(nC23nC25nC29nC31).Sedges and ferns are prolific in many tropical ecosystems (14). These plants both have variable and therefore nondiagnostic n-alkane profiles. However, sedges produce distinctive phenolic compounds [e.g., 5-n-tricosylresorcinol (nR23)] and ferns produce distinctive midchain diols [e.g., 1,13-dotriacontanediol (C32-diol)] (SI Discussion).Lignin monomers provide evidence for woody and nonwoody plants. This refractory biopolymer occurs in both leaves and wood, serves as a structural tissue, and accounts for up to half of the total organic carbon in modern vegetation (11). Lignin is composed of three phenolic monomer types that show distinctive distributions in woody and herbaceous plant tissues. Woody tissues from dicotyledonous trees and shrubs contain syringyl (S) and vanillyl (V) phenols (12), whereas cinnamyl (C) phenols are exclusively found in herbaceous tissues (12). The relative abundance of C versus V phenols (C/V) is widely used to distinguish between woody and herbaceous inputs to sedimentary and soil organic matter (15).Plant biomarker 13C/12C ratios (expressed as δ13C values) are sensitive indicators of community composition, ecosystem structure, and climate conditions (8). Most woody plants and forbs in eastern Africa use C3 photosynthesis (6), whereas arid-adapted grasses use C4 photosynthesis (8, 14). These two pathways discriminate differently against 13C during photosynthesis, resulting in characteristic δ13C values for leaf waxes derived from C3 (about –36.0‰) and C4 (–21.0‰) plants (16). Carbon isotopic abundances of phenolic monomers of lignin amplify the C3–C4 difference and range between ca. –34.0‰ (C3) and –14.0‰ (C4) in tropical ecosystems (15). Terrestrial C3 plant δ13C values decrease with increased exposure to water, respired CO2, and shade (8), with lowest values observed in moist regions with dense canopy (17). Although concentration and δ13C values of atmospheric CO2 can affect C3 plant δ13C values (17), this influence is not relevant to our work here, which focuses on a single time window (SI Discussion). The large differences in leaf-wax δ13C values between closed C3 forest to open C4 grassland are consistent with soil organic carbon isotope gradients across canopy-shaded ground surfaces (6) and serve as a quantitative proxy for woody cover (fwoody) in savannas (8).As is observed for nonhuman primates, hominin dietary choices were likely shaped by ecosystem characteristics over habitat scales of 1–1,000 m2 (35). To evaluate plant distributions at this small spatial scale (9), we excavated 71 paleosol samples from close-correlated trenches across a ∼25,000-m2 area that included FLK Zinjanthropus Level 22 (FLK Zinj) at Olduvai Gorge (Fig. 1). Recent excavations (1821) at multiple trenches at four sites (FLKNN, FLKN, FLK, and FLKS, Fig. 1D) exposed a traceable thin (5–50 cm), waxy green to olive-brown clay horizon developed by pedogenic alterations of playa lake margin alluvium (22). Weak stratification and irregular redox stains suggest initial soil development occurred during playa lake regression (18, 22), around 1.848 Ma (ref. 23 and SI Discussion). To date, craniodental remains from at least three hominin individuals (1820), including preadolescent early Homo and Paranthropus boisei, were recovered from FLK Zinj. Fossils and artifacts embedded in the paleosol horizon often protrude into an overlying airfall tuff (18, 19), which suggests fossil remains were catastrophically buried in situ under volcanic ash. Rapid burial likely fostered the exceptional preservation of both macrofossils (10) and plant biomarkers across the FLK Zinj land surface.Open in a separate windowFig. 1.Location and map of FLK Zinj paleosol excavations. (A and B) Location of FLK Zinj as referenced to reconstructed depositional environments at Olduvai Gorge during the early Pleistocene (18, 22) and the modern gorge walls. The perennial lake contained shallow saline–alkaline waters that frequently flooded the surrounding playa margin (i.e., floodplain) flats. (C) Outline of FLK Zinj paleosol excavation sites used for our spatial biomarker reconstructions. (D) Concentric (5 m) gridded distribution map of FLK Zinj paleosol excavations relative to previous archaeological trenches (1821). Major aggregate complexes (FLKNN, FLKN, FLK, and FLKS) are color-coded to show excavation-site associations.Plant biomarker signatures reveal distinct types of vegetation juxtaposed across the FLK Zinj land surface (Figs. 24 and Fig. S1). In the northwest, FLKNN trenches show high nC23 δ13C values (Fig. 2B) as well as high C/V and Paq values (Figs. 3 and and4A).4A). They indicate floating or submerged aquatic plants (macrophytes) in standing freshwater (13), a finding that is consistent with nearby low-temperature freshwater carbonates (tufa), interpreted to be deposited from spring waters (22). Adjacent FLKN trenches have lower Paq values (Fig. 4A) with occurrences of fern-derived C32-diol and sedge-derived nR23 (Fig. 2 C and D). These biomarker distributions indicate an abrupt (around 10 m) transition from aquatic to wetland vegetation. Less than 100 m away (Fig. 1C), low nC31 δ13C values (Fig. 2A) and low C/V and very low Paq values (Figs. 3 and and4A)4A) collectively indicate dense woody cover (Fig. 4B). In the farthest southeastern (FLKS) trenches, high C/V values and high δ13C values for C lignin phenols (Fig. 3) indicate open C4 grassland.Open in a separate windowFig. 2.Spatial distributions and δ13C values for plant biomarkers across FLK Zinj. Measured and modeled δ13C values (large and smaller circles, respectively) are shown for (A) nC31 from terrestrial plants, (B) nC23 from (semi)aquatic plants, (C) C32-diol from ferns, and (D) nR23 from sedges (see refs. 12 and 13 and SI Discussion). Modeled values [inverse distance-weighted (9)] account for spatial autocorrelation (15-m radius) in standing biomass (35) over scales of soil organic matter accumulation (11). Black dots represent paleosols with insufficient plant biomarker concentrations for isotopic analysis.Open in a separate windowFig. 3.Molecular and isotopic signatures for lignin phenols across FLK Zinj. Bivariate plots are shown for diagnostic lignin compositional parameters (see refs. 12 and 15 and Fig. 1C). Symbols are colored according to respective δ13C values for the C lignin phenol, p-coumaric acid. FLK symbols are uncolored due to insufficient p-coumaric acid concentrations for isotopic analysis. Representative lignin compositional parameters (12, 15) are shown for monocotyledonous herbaceous tissues (G), dicotyledonous herbaceous tissues (H), cryptogams (N), and dicotyledonous woody tissues (W).Open in a separate windowFig. 4.Spatial relationships shared between local plant resources and hominin remains. Measured and modeled values (large and smaller circles, respectively) are shown for (A) Paq (13) and (B) fwoody (8). Modeled values [inverse distance-weighted (9)] account for spatial autocorrelation (15-m radius) in standing biomass (35) over scales of soil organic matter accumulation (11). (C) Kernel density map of cut-marked bones (1821) across the FLK Zinj land surface (Fig. S4). High estimator values indicate hotspots of hominin butchery (Fig. S5). A shaded rectangle captures the area (ca. 0.68 probability mass) with highest cut-marked bone densities and is shown in A and B for reference.Open in a separate windowFig. S1.Total ion chromatograms for saturated hydrocarbons in representative paleosols at (A) FLKNN, (B) FLKN, (C) FLK, and (D) FLKS. C23, tricosane; C25, pentacosane; C29 nonacosane; C31, hentriacontane.Biomarkers define a heterogeneous landscape at Olduvai and suggest an influence of local resources on hominin diets and behavior. It is recognized (2, 2426) that early Homo species and P. boisei had similar physiological characteristics. These similarities in physical attributes suggest behavioral differences were what allowed for overlapping ranges and local coexistence (sympatry) of both hominins. For instance, differences in seasonal subsistence strategies or different behavior during periods of drought and limited food could have reduced local hominin competition and fostered diversification via niche specialization (2729).Physical and isotopic properties of fossil teeth indicate P. boisei was more water-dependent [low enamel δ18O values (24)] and consumed larger quantities of abrasive, 13C-enriched foodstuffs [flat-worn surfaces (25) and high enamel δ13C values (26)] than coexisting early Homo species. Although 13C-enriched enamel is commonly attributed to consumption of C4 grasses or meat from grazers (14), this was not likely, because P. boisei craniodental features are inconsistent with contemporary gramnivores (24, 25) or extensive uncooked flesh mastication (26). Numerous scholars have proposed the nutritious underground storage organs (USOs) of C4 sedges were a staple of hominin diets (14, 24, 26, 27). Consistent with this suggestion, occurrences of nR23 attest to the presence of sedges at FLKNN and FLKN (Fig. 2D). However, the low δ13C values measured for nR23 at these same sites (Fig. 2D and Fig. S2) indicate C3 photosynthesis (12, 16), a trait common in modern sedges that grow in alkaline wetlands and lakes (30) (Fig. S3). Thus, biomarker signatures support the presence of C3 sedges in the wetland area of FLK Zinj.Open in a separate windowFig. S2.Total ion chromatogram [TIC (A)] and selected ion chromatograms for derivatized 5-n-alkylresorcinols [m/z 268 (●)] and midchain diols [m/z 369 (○)] from a representative paleosol at FLKN. Also shown are δ13C values for homologous (B) 5-n-alkylresorcinols and (C) midchain diols. C32-diol, dotriacontanediol; nR23, tricosylresorcinol.Open in a separate windowFig. S3.Summary phyogenetic consensus tree of Cyperaceae (sedges) based on nucleotide (rcbL and ETS1f) sequence data (5054, 95, 96). Important taxonomic distinctions discussed in SI Discussion, Fern Alkyldiols are shown explicitly. Triangle-enclosed digits represent the number of additional branches at different levels of taxonomic classification. CEFA, Cypereae Eleocharideae Fuireneae Abildgaardieae; CSD, Cariceae Scirpeae Dulichieae.Alternative foodstuffs with abrasive, 13C-enriched biomass include seedless vascular plants (cryptogams), such as ferns and lycophytes [e.g., quillworts (27, 30)]. Ferns are widely distributed throughout eastern Africa in moist and shaded microhabitats (31) and are often found near dependable sources of drinking water (32). Today, ferns serve as a dietary resource for humans and nonhuman primates alike (27), and fiddlehead consumption is consistent with the inferred digestive physiology [salivary proteins (33)] and the microwear on molars (34) of P. boisei in eastern Africa (25, 26). Ferns were present at FLKN, based on measurements of C32-diol (Fig. 2D). Further, the high δ13C values measured for these compounds are consistent with significant fern consumption by P. boisei at Olduvai Gorge.Ferns and grasses were not the only plant foods present during the time window documented by FLK Zinj. Further, the exclusive reliance on a couple of dietary resources was improbable for P. boisei, because its fossils occur in diverse localities (2426). Aquatic plants are an additional candidate substrate, as evidenced by high Paq values at FLKNN and FLKN (Fig. 4A). Floating and submerged plants proliferate in wetlands throughout eastern Africa today (13, 14), and many produce nutritious leaves and rootstock all year long (27, 28). Although C4 photosynthesis is rare among modern macrophytes (30), they can assimilate bicarbonate under alkaline conditions, which results in C4-like isotope signatures in their biomass (30). Their leaf waxes, such as nC23 (13), are both present and carry 13C-enriched signatures at FLKNN and FLKN (Fig. 2B). It is also likely that aquatic macrophytes sustained invertebrates and fish with comparably 13C-enriched biomass, as they do in modern systems (14), and we suggest aquatic animal foods could have been important in P. boisei diets (27, 28).Biomarkers across the FLK Zinj soil horizon resolve clear patterns in the distribution of plants and water and suggest critical resources that shaped hominin existence at Olduvai Gorge. The behavioral implications of local conditions require understanding of regional climate and biogeography (35, 7), because hominin species likely had home ranges much larger than the extent of excavated sites at FLK Zinj. Lake sediments at Olduvai Gorge include numerous stacked tuffs with precise radiometric age constraints (23). These tephrostratigraphic correlations (21) tie the FLK Zinj landscape horizon to published records of plant biomarkers in lake sediments that record climate cycles and catchment-scale variations in ecology. Correlative lake sediment data indicate the wet and wooded microhabitats of FLK Zinj sat within a catchment dominated by arid C4 grassland (8). Under similarly arid conditions today, only a small fraction of landscape area (ca. 0.05) occurs within 5 km of either forest or standing freshwater (35). Given a paucity of shaded refuge and potable water in the catchment, the concentration of hominin butchery debris (1821) exclusively within the forest microhabitat and adjacent to a freshwater wetland (Fig. 4) is notable. We suggest the spatial patterns defined by both macro- and molecular fossils reflect hominins engaged in social transport of resources (15), such as bringing animal carcasses and freshwater-sourced foods from surrounding grassy or wetland habitats to a wooded patch that provided both physical protection and access to water.  相似文献   

18.
A simple electrochemically mediated method for the conversion of alkyl carboxylic acids to their borylated congeners is presented. This protocol features an undivided cell setup with inexpensive carbon-based electrodes and exhibits a broad substrate scope and scalability in both flow and batch reactors. The use of this method in challenging contexts is exemplified with a modular formal synthesis of jawsamycin, a natural product harboring five cyclopropane rings.

Boronic acids are among the most malleable functional groups in organic chemistry as they can be converted into almost any other functionality (13). Aside from these versatile interconversions, their use in the pharmaceutical industry is gaining traction, resulting in approved drugs such as Velcade, Ninlaro, and Vabomere (4). It has been shown that boronic acids can be rapidly installed from simple alkyl halides (519) or alkyl carboxylic acids through the intermediacy of redox-active esters (RAEs) (Fig. 1A) (2024). Our laboratory has shown that both Ni (20) and Cu (21) can facilitate this reaction. Conversely, Aggarwal and coworkers (22) and Li and coworkers (23) demonstrated photochemical variations of the same transformation. While these state-of-the-art approaches provide complementary access to alkyl boronic acids, each one poses certain challenges. For example, the requirement of excess boron source and pyrophoric MeLi under Ni catalysis is not ideal. Although more cost-effective and operationally simple, Cu-catalyzed borylation conditions can be challenging on scale due to the heterogeneity resulting from the large excess of LiOH•H2O required. In addition to its limited scope, Li and coworkers’ protocol requires 4 equivalence of B2pin2 and an expensive Ir photocatalyst. The simplicity of Aggarwal and coworkers’ approach is appealing in this regard and represents an important precedent for the current study.Open in a separate windowFig. 1.(A) Prior approaches to access alkyl boronic esters from activated acids. (B) Inspiration for initiating SET events electrochemically to achieve borylation. (C) Summary of this work.At the heart of each method described above, the underlying mechanism relies on a single electron transfer (SET) event to promote decarboxylation and form an alkyl radical species. In parallel, the related borylation of aryl halides via a highly reactive aryl radical can also be promoted by SET. While numerous methods have demonstrated that light can trigger this mechanism (Fig. 1B) (16, 2531), simple electrochemical cathodic reduction can elicit the same outcome (3235). It was postulated that similar electrochemically driven reactivity could be translated to alkyl RAEs. The development of such a transformation would be highly enabling, as synthetic organic electrochemistry allows the direct addition or removal of electrons to a reaction, representing an incredibly efficient way to forge new bonds (3640). This disclosure reports a mild, scalable, and operationally simple electrochemical decarboxylative borylation (Fig. 1C) not reliant on transition metals or stoichiometric reductants. In addition to mechanistic studies of this interesting transformation, applications to a variety of alkyl RAEs, comparison to known decarboxylative borylation methods, and a formal synthesis of the polycyclopropane natural product jawsamycin [(–)-FR-900848] are presented.  相似文献   

19.
An internal ribosome entry site (IRES) initiates protein synthesis in RNA viruses, including the hepatitis C virus (HCV). We have discovered ligand-responsive conformational switches in viral IRES elements. Modular RNA motifs of greatly distinct sequence and local secondary structure have been found to serve as functionally conserved switches involved in viral IRES-driven translation and may be captured by identical cognate ligands. The RNA motifs described here constitute a new paradigm for ligand-captured switches that differ from metabolite-sensing riboswitches with regard to their small size, as well as the intrinsic stability and structural definition of the constitutive conformational states. These viral RNA modules represent the simplest form of ligand-responsive mechanical switches in nucleic acids.Internal ribosome entry site (IRES) elements provide an alternative mechanism for translation initiation by directing the assembly of functional ribosomes directly at the start codon in a process that does not require 5′ cap recognition or ribosomal scanning and that is independent of many host initiation factors (14). The genomes of Flaviviridae and Picornaviridae contain elements that share similarity with the archetypical hepatitis C virus (HCV) IRES in overall domain organization, but not sequence or details of secondary structure (5). The HCV IRES adopts a complex architecture of four independently folding domains (Fig. 1A) (6). Domain II is nearly 100% conserved in clinical isolates (7) and has analogous counterparts in other viral IRES elements, all of which display some secondary structure similarity, but significant sequence variation in the subdomain IIa-like internal loop (Fig. 1B). Domain II has been shown to promote stable entry of HCV and classic swine fever virus (CSFV) mRNA at the decoding groove of the 40S subunit (810) and is required for initiation factor removal before ribosomal subunit joining (11), as well as adjustment of initiator tRNA orientation (12). The transition from initiation to elongation stages of translation depends critically on domain II (13). Recently, direct interaction of HCV domain II with initiator tRNA has been demonstrated (14). In HCV, subdomain IIa folds into an L-shaped motif (15) (Fig. 1C) that introduces a 90° bend in domain II (16) and directs the IIb hairpin toward the E-site at the ribosomal subunit interface (17, 18).Open in a separate windowFig. 1.Structures and ligands of viral IRES. (A) The IRES in the 5′ UTR of the HCV genome. The location of subdomain IIa is highlighted by an orange box. The viral genome encodes structural (S) and nonstructural (NS) proteins and contains a structured 3′ UTR. (B) Secondary structure predictions of domain II motifs in viral IRES elements from HCV and other flaviviruses, including CSFV and BVDV, as well as picornaviruses such as AEV and SVV. Non-Watson–Crick base pairs are indicated by the ○ symbol. Sequence conservation is indicated in red. (C) Crystal structure of the subdomain IIa RNA from HCV. (D) Benzimidazole (1) and diaminopiperidine (2) inhibitors of IRES-driven translation that target the HCV subdomain IIa. (E) Crystal structure of the HCV subdomain IIa RNA in complex with inhibitor 1.The HCV IRES subdomain IIa is the target for viral translation inhibitors (Fig. 1D) that bind to the internal loop and block translation by capturing distinct conformational states of the RNA (7). Structure analysis revealed that benzimidazole inhibitors such as compound 1 (19, 20) interact with an extended architecture of IIa in which the stems flanking the internal loop are coaxially stacked on both sides of the ligand-binding pocket (Fig. 1E) (21). In contrast, diaminopiperidine compounds such as 2 bind and lock the IIa RNA in a bent conformation that corresponds to the ligand-free state (22). Conformational capture of the subdomain IIa switch by ligands in solution was demonstrated by FRET experiments and established as a mechanism of IRES inhibition (23). On the basis of these findings, it was proposed that subdomain IIa may be the target for a cognate biological ligand whose adaptive recognition by the RNA motif may facilitate ribosome release from the IRES-bound complex (7).Here, we have explored potential candidates for a cognate ligand of the subdomain IIa switch and investigated the structural and functional conservation of similar ligand responsive switch motifs in other IRES RNAs.  相似文献   

20.
Enzymes use protein architectures to create highly specialized structural motifs that can greatly enhance the rates of complex chemical transformations. Here, we use experiments, combined with ab initio simulations that exactly include nuclear quantum effects, to show that a triad of strongly hydrogen-bonded tyrosine residues within the active site of the enzyme ketosteroid isomerase (KSI) facilitates quantum proton delocalization. This delocalization dramatically stabilizes the deprotonation of an active-site tyrosine residue, resulting in a very large isotope effect on its acidity. When an intermediate analog is docked, it is incorporated into the hydrogen-bond network, giving rise to extended quantum proton delocalization in the active site. These results shed light on the role of nuclear quantum effects in the hydrogen-bond network that stabilizes the reactive intermediate of KSI, and the behavior of protons in biological systems containing strong hydrogen bonds.Although many biological processes can be well-described with classical mechanics, there has been much interest and debate as to the role of quantum effects in biological systems ranging from photosynthetic energy transfer, to photoinduced isomerization in the vision cycle and avian magnetoreception (1). For example, nuclear quantum effects, such as tunneling and zero-point energy (ZPE), have been observed to lead to kinetic isotope effects of greater than 100 in biological proton and proton-coupled electron transfer processes (2, 3). However, the role of nuclear quantum effects in determining the ground-state thermodynamic properties of biological systems, which manifest as equilibrium isotope effects, has gained significantly less attention (4).Ketosteroid isomerase (KSI) possesses one of the highest enzyme unimolecular rate constants and thus, is considered a paradigm of proton transfer catalysis in enzymology (511). The remarkable rate of KSI is intimately connected to the formation of a hydrogen-bond network in its active site (Fig. 1A), which acts to stabilize a charged dienolate intermediate, lowering its free energy by ∼11 kcal/mol (1 kcal = 4.18 kJ) relative to solution (Fig. S1) (6). This extended hydrogen-bond network in the active site links the substrate to Asp103 and Tyr16, with the latter further hydrogen-bonded to Tyr57 and Tyr32, which is shown in Fig. 1A.Open in a separate windowFig. 1.KSI⋅intermediate and KSID40N ? inhibitor complex. Schematic depiction of (A) the KSI⋅intermediate complex during the catalytic cycle (Fig. S1) and (B) a complex between KSID40N and phenol, an inhibitor that acts as an intermediate analog. Both the intermediate and the inhibitor are stabilized by a hydrogen-bond network in the active site of KSI. (C) Image of KSID40N with the tyrosine triad enlarged and the atoms O16, H16, O32, H32, and O57 labeled (shown with Tyr57 deprotonated) (16).The mutant KSID40N preserves the structure of the wild-type (WT) enzyme while mimicking the protonation state of residue 40 in the intermediate complex (Fig. 1B), therefore permitting experimental investigation of an intermediate-like state of the enzyme (6, 1214). Experiments have identified that, in the absence of an inhibitor, one of the residues in the active site of KSID40N is deprotonated (12). Although one might expect the carboxylic acid of Asp103 to be deprotonated, the combination of recent 13C NMR and ultraviolet visible spectroscopy (UV-Vis) experiments has shown that the ionization resides primarily on the hydroxyl group of Tyr57, which possesses an anomalously low pKa of 6.3 ± 0.1 (12). Such a large tyrosine acidity is often associated with specific stabilizing electrostatic interactions (such as a metal ion or cationic residue in close proximity), which is not the case here, suggesting that an additional stabilization mechanism is at play (15).One possible explanation is suggested by the close proximity of the oxygen (O) atoms on the side chains of the adjacent residues Tyr16 (O16) and Tyr32 (O32) to the deprotonated O on Tyr57 (O57) (Fig. 1C) (16). In several high-resolution crystal structures, these distances are found to be around 2.6 Å (14, 16, 17), which is much shorter than those observed in hydrogen-bonded liquids such as water, where O–O distances are typically around 2.85 Å. Such short heavy-atom distances are only slightly larger than those typically associated with low-barrier hydrogen bonds (1820), where extensive proton sharing is expected to occur between the atoms. In addition, at these short distances, the proton’s position uncertainty (de Broglie wavelength) becomes comparable with the O–O distance, indicating that nuclear quantum effects could play an important role in stabilizing the deprotonated residue (Fig. 1C). In this work, we show how nuclear quantum effects determine the properties of protons in the active-site hydrogen-bond network of KSID40N in the absence and presence of an intermediate analog by combining ab initio path integral simulations and isotope effect experiments.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号