Category: 7. Science

  • How old is Jupiter: Meteorite ‘raindrops’ help scientists pin down gas giant’s age

    How old is Jupiter: Meteorite ‘raindrops’ help scientists pin down gas giant’s age

    In some cases, to learn more about our solar system, all we have to do is look at evidence found on Earth.

    Researchers from Japan’s Nagoya University and the Italian National Institute for Astrophysics (INAF) have determined how mysterious molten “raindrops” in meteorites formed — and used that information to date Jupiter’s formation.

    Continue Reading

  • AUF1 protein shown to suppress cellular senescence and altered metabolism

    AUF1 protein shown to suppress cellular senescence and altered metabolism

    A new research paper was published in Volume 17, Issue 7 of Aging (Aging-US) on July 24, 2025, titled “RNA-binding protein AUF1 suppresses cellular senescence and glycolysis by targeting PDP2 and PGAM1 mRNAs.”

    In this study, Hyejin Mun, Chang Hoon Shin, Mercy Kim, Jeong Ho Chang, and Je-Hyun Yoon from the University of Oklahoma and Kyungpook National University investigated how changes in cellular metabolism contribute to aging. Their findings offer potential targets for therapies aimed at slowing or reducing the effects of aging.

    As cells age, they often lose their ability to divide and begin releasing harmful signals that damage nearby tissues. This process, called cellular senescence, is linked to many age-related diseases. A key feature of senescent cells is their altered metabolism, where they use more glucose and oxygen, even when oxygen levels are low. This leads to the production of inflammatory substances and fatty acids, which can accelerate tissue damage. The study examined how these metabolic changes are controlled at the molecular level.

    Researchers found that AUF1, a protein that binds to RNA, normally helps prevent aging by breaking down two enzymes involved in glucose metabolism: PGAM1 and PDP2. When AUF1 is missing or inactive, these enzymes build up. This causes the cell to produce more energy and inflammatory molecules, which are common features of senescent cells.

    “Our high throughput profiling of mRNAs and proteins from Human Diploid Fibroblasts (HDFs) revealed that the expression of pyruvate metabolic enzymes is inhibited by the anti-senescent RNA-binding protein (RBP) AUF1 (AU-binding Factor 1).”

    The team also identified another protein, MST1, which becomes active during cellular stress and aging. MST1 modifies AUF1 in a way that stops it from doing its protective job. As a result, PGAM1 and PDP2 accumulate, leading to faster aging of the cell. Experiments using human fibroblast cells and mouse models confirmed that higher levels of these enzymes are linked to stronger signs of cellular aging.

    These findings improve our understanding of how metabolism affects the aging process. They highlight the MST1-AUF1-PDP2/PGAM1 pathway as a key factor in the metabolic shift seen in aging cells. Since these enzymes and proteins are already known to be involved in other diseases, existing or future therapies might be used to block this pathway and reduce the effects of aging.

    This study offers a new direction for senotherapy-a field focused on treating or removing aging cells. By adjusting glucose metabolism through AUF1 and its targets, scientists believe it may be possible to slow aging or limit its effects on tissue function. More research is needed, but these insights could lead to new strategies for managing age-related diseases and promoting healthier aging.

    Source:

    Journal reference:

    Mun, H., et al. (2025). RNA-binding protein AUF1 suppresses cellular senescence and glycolysis by targeting PDP2 and PGAM1 mRNAs. Aging. doi.org/10.18632/aging.206286

    Continue Reading

  • Older Species Have the Largest Ranges

    Older Species Have the Largest Ranges


    Register for free to listen to this article

    Thank you. Listen to this article using the player above.


    Want to listen to this article for FREE?


    Complete the form below to unlock access to ALL audio articles.

    Every living species on Earth has a unique geographical range, with some being widespread and others being very narrow. Several factors shape a species’ range size – and one of them is the evolutionary age of a species. To investigate how evolutionary age is related to present-day range size, a research team led by scientists from the German Centre for Integrative Biodiversity Research (iDiv), Leipzig University and Naturalis Biodiversity Center compared over 26,000 species of mammals, birds, reptiles, amphibians, reef fishes, and palms.

    More than 40,000 species are facing extinction worldwide. Species with narrow geographical ranges are known to face a higher risk of extinction compared to widespread species: They tend to have lower overall abundance and smaller local populations, making them vulnerable to environmental perturbations. Although range size is one of the strongest determinants of extinction risk, the causes underlying the wide variation in natural range sizes remain poorly understood.

    An international team of researchers led by Dr Adriana Alzate, alumna of iDiv and Leipzig University, compared data on the evolutionary age and range sizes of over 26,000 species from seven major taxonomic groups: birds, reptiles, amphibians, reef fishes, palms, and terrestrial as well as marine mammals.

    More time, larger range?

    The researchers show that, on average, older species have larger ranges across all groups except for marine mammals – a finding which did not come as a big surprise, but had not been proven so far. “Older species are expected to have larger distributions because they have had more time, sometimes several millions of years, to expand their ranges since first appearing,” explains first author Adriana Alzate, Guest Researcher at Naturalis Biodiversity Center. “Over evolutionary timescales, these species have had more opportunities to reproduce, disperse, colonize and adapt to diverse environments, allowing them to occupy broader geographical areas.”

    Good dispersers quickly attain large ranges

    But it is not only a species’ age that influences its range size. Some species are good dispersers and can move easily across barriers or over great distances, such as birds with long, narrow wings with pointed tips, and palms with large fruits, that are dispersed by wide-ranging large-bodied vertebrates. These species may attain large ranges faster than less dispersive species. Therefore, good dispersers might have larger ranges than expected based on age only. By contrast, the study shows that the effect of species age is likely more pronounced on less dispersive species, such as amphibians.

    Geographical context also plays a major role. On islands, the maximum range size that native species can attain is geographically constrained. The new research confirmed that island-restricted species have smaller range sizes than species that are not restricted to islands – but it also brought to light an unexpected relationship: On islands, range size differences between young and old species are greater than on the mainland. Dr Roberto Rozzi, Curator of Palaeontology at the Central Repository of Natural Science Collections of Martin Luther University Halle-Wittenberg and iDiv alumnus: “Island dynamics and ontogeny modulate the relationship between age and range size. Release from predators and competitors may have enabled early island colonizers, typically ecological generalists, to achieve broader ranges than expected based on age only.”

    Environmental change puts narrow-ranged species at risk

    The smaller a species’ range, the higher the risk for it to go extinct. Understanding the dynamics that shape a species’ range size is crucial for predicting its vulnerability to extinction and tailoring conservation efforts to local conditions and needs.  “This is even more important in the context of changing environmental conditions, because not all species may be able to keep up with these changes”, says senior author Dr Renske Onstein, junior group leader at iDiv and group leader “Biodiversity Hotspots” at Naturalis Biodiversity Center. “Possibly, older species have the genetic makeup to more readily adapt and therefore persist in their relatively large ranges. This needs further testing with genetic data, for example, providing exciting possibilities for future research”.

     

    Reference: Alzate A, Rozzi R, Velasco JA, et al. Evolutionary age correlates with range size across plants and animals. Nat Commun. 2025. doi:10.1038/s41467-025-62124-y


    This article has been republished from the following materials. Note: material may have been edited for length and content. For further information, please contact the cited source. Our press release publishing policy can be accessed here.

    Continue Reading

  • ‘Landmark moment’: Scientists reveal the most detailed photo of a solar flare ever captured

    ‘Landmark moment’: Scientists reveal the most detailed photo of a solar flare ever captured

    The Inouye Solar Telescope captured this image of a solar flare on August 8, 2024. (Image credit: NSF/NSO/AURA, CC-BY)

    The world’s largest solar telescope just captured the highest-resolution images of a solar flare to date — and they’re spectacular.

    Researchers trained the Hawaii-based Daniel K. Inouye Solar Telescope on the final stages of a powerful X-class solar flare on Aug. 8, 2024, capturing detailed images of chaotic loops of plasma at the sun’s surface. The observations could help scientists understand the mechanics of solar flares and improve predictions of future flares.

    Continue Reading

  • Glow-in-the-dark houseplants shine in rainbow of colours

    Glow-in-the-dark houseplants shine in rainbow of colours

    University students might soon have something other than black-light posters to brighten their dorm rooms. Researchers have created glow-in-the-dark plants by injecting succulents with materials similar to those that make the posters light up. The fleshy plants shine as brightly as a night light, and can be made to do so in a wide variety of colours — a first for glowing houseplants, according to the team.

    The researchers, led by Xuejie Zhang, a materials scientist at the South China Agricultural University in Guangzhou, describe today how they produced the plants in the journal Matter1. They have applied for a patent on the technology, which they hope will lead to decorative installations and living lighting.

    The idea of making glowing plants has captivated scientists since the late 1980s, when researchers made the first bioluminescent plant2 by inserting a gene from a firefly (Photinus pyralis) into a type of tobacco (Nicotiana tabacum). This work laid the foundation for the first genetically engineered luminescent houseplant to come on the market in the United States, last year. The biotechnology firm Light Bio in Sun Valley, Idaho, sells the petunia (Petunia hybrida), which glows a very faint green thanks to genes from a light-emitting mushroom.

    Leafy greens … and blues and reds

    Unlike the petunia, which emits light through chemical reactions in its cells, the succulent glows because of materials injected into its leaves. These materials — phosphor particles made of strontium and aluminium dosed with other metals — absorb energy from light at one wavelength, store some of that energy and then slowly re-emit it at a different wavelength for several hours. For instance, one material the scientists injected into their succulents absorbs ultraviolet and blue light, and re-emits it as green light.

    This type of ‘afterglow’ phosphor is used in glow-in-the-dark toys and paints, and as an imaging tracer for laboratory animals. Whereas genetically engineered bioluminescent plants are, so far, limited in the range of colours they emit, afterglow phosphors span a wide variety of hues, including red and blue, and they can be combined to produce a white glow.

    Video of succulents being infused with afterglow phosphor particles that absorb and slowly release light.

    This video shows how the tiny phosphor particles spread through the succulent leaves after they are injected.Credit: Liu et al./Matter

    The researchers purchased phosphors containing strontium aluminate and ground them down to particles of various sizes before injecting them into an assortment of plants. They found that particles around 7 micrometres in diameter glowed more brightly than did nanoparticles in plants, and were able to fill up the interior tissues of succulent leaves for a stronger, more uniform glow. By contrast, plants with simple leaf structures, such as tobacco plants and pak choi, emitted a more patchy glow.

    The plant favoured by the team is the succulent Echevaria ‘Mebina’, a common houseplant that grows rosettes of dense, fleshy leaves. To make every leaf glow, the researchers had to inject each one with phoshor particles, a process that takes about ten minutes. The luminescence — which the team generated in hues of blue-green, blue-violet, green, red and white — lasted as long as 120 minutes after exposing the plant to tailored wavelengths of light or sunlight, and could be triggered again and again over the 10 days of the study.

    Continue Reading

  • Attosecond control and measurement of chiral photoionization dynamics

    Attosecond control and measurement of chiral photoionization dynamics

    Chirality is traditionally viewed as a structural property of matter. The spatial arrangement of atoms in molecules and materials indeed defines their handedness. In a simplified view, structural chirality is often used to explain chiral recognition. However, recent research indicates that structural chirality is insufficient to fully understand chiral phenomena. In particular, growing evidence points to a functional role of chiral electron dynamics in enabling and mediating chiral interactions at the electronic level5. Such dynamics are also thought to play a role in advanced applications, including spintronics6, unidirectional molecular machines7 and chiral-sensitive biosensing8. Furthermore, a wide range of chiroptical techniques, such as electronic circular dichroism9 (CD), PECD10,11 and chiral high-harmonic generation12,13, are now understood to examine the underlying chiral dynamics of electrons rather than merely structural asymmetries. These developments motivate a broader perspective on chirality that encompasses not only static structure but also the dynamic behaviour of electrons in chiral systems.

    In spite of the fundamental importance of chiral electron dynamics, spectroscopy based on attosecond pulses has so far been lacking the crucial capability of chiral discrimination. The key obstacle for expanding attosecond science into the realms of chiral molecules and materials has been the lack of circularly polarized attosecond light pulses. As a consequence, all pioneering experiments in this field have relied on femtosecond pulses, for example, refs. 12,13,14,15,16,17,18,19,20,21. Notably, strong-field ionization with intense two-colour femtosecond laser pulses has been used to measure phase shifts17 and control the asymmetry of photoelectron angular distributions (PADs) in the strong-field ionization of chiral molecules20,22. Recently, some of the present authors have developed attosecond metrology in circular polarization by introducing a plug-in apparatus for the generation and demonstration of a general methodology for the complete characterization of circularly polarized attosecond pulses1. This new capability has recently been applied to study the photoionization dynamics of atoms2,23.

    Here we introduce attosecond chiroptical spectroscopy by reporting the first application of circularly polarized attosecond pulses to chiral molecules. This development, combined with momentum-vector-resolved electron-ion-coincidence spectroscopy, allows us to measure and control chiral electron dynamics on their natural attosecond timescale. Specifically, our study demonstrates attosecond coherent control over PECD based on the constructive or destructive interference between pairs of well-defined photoionization pathways. It also reveals the characteristic dependence of chiral photoionization delays on both the polar and the azimuthal angles of photoemission in the light-propagation frame.

    The experimental set-up is illustrated in Fig. 1a, with details given in Methods. The circularly polarized extreme-ultraviolet (XUV) attosecond pulse train (APT) of a single, common helicity was produced by the non-collinear HHG scheme using a compact plug-in apparatus1. A weak linearly or circularly polarized (co-rotating) near-IR (800 nm) laser pulse was spatio-temporally overlapped with the XUV APT with attosecond temporal stability to induce XUV-IR two-photon transitions using the reconstruction of attosecond beating by interference of two-photon transitions (RABBIT) approach. In this study, we define the light-propagation direction to be the z axis and the light polarization to be in the xy plane, so that the CD manifests itself in the angular distribution with respect to the angle θ = arccos(pz/ptotal), in which ptotal and pz are the total momentum of the electron and its z component, respectively. The azimuthal angle in the polarization plane ϕ = atan(px/py) was integrated over from −15° to +15°. Enantiomerically pure samples of methyloxirane (MeOx, also known as propylene oxide, C3H6O, ≥99.5% ee) were delivered into a cold-target recoil-ion-momentum spectroscopy (COLTRIMS) set-up24,25 by supersonic expansion through the nozzle with an aperture of 30 µm. The 3D momenta of photoelectrons and ionic fragments were measured in coincidence in the ultrahigh-vacuum chamber (about 10−10 mbar).

    Fig. 1: Attosecond chiroptical coincidence spectroscopy.

    a, Experimental set-up. The ionic fragments and emitted photoelectrons are measured in coincidence using COLTRIMS. b, Measured 2D ionic spectrum as a function of the mass-over-charge ratio and the position on the ion detector along the molecular-beam direction in the presence of the XUV field only. c, The projection of the distribution in b on the mass-over-charge axis. Note that the counts are shown on a logarithmic scale. The two small peaks above m/q = 58 correspond to 13C isotopomers of the parent ion. d,e, Measured photoelectron spectra in coincidence with the intact parent ion (d) and the C2H4+ fragment ion (e). In d, the observation of a series of discrete equidistant peaks confirms that photoionization to the ({mathop{X}limits^{ sim }}^{+}) ground state of the cation leaves the molecule intact. In e, the experimental data are fitted by two sets of Gaussian peaks corresponding to the ({widetilde{A}}^{+}) and ({widetilde{B}}^{+}) ionic states, respectively. The ionization energies of the relevant states are taken from the ref. 26. a.u., arbitrary units.

    Figure 1b shows the ionic fragment distribution as a function of mass-over-charge ratio (m/q) and detector position and Fig. 1c presents the projection onto the m/q axis, both recorded using attosecond XUV pulses only. After photoionization by the attosecond pulses comprising harmonic 7 (H7) to harmonic 13 (H13), we observe the undissociated parent ion (C3H6O+) and two broader distributions assigned to CH3+ and C2H4+ fragments, in agreement with previous synchrotron studies26. The photoelectron spectrum identifies the electronic state of the ion that was accessed in the ionization step, whereas the ionic species reveal whether and how a given electronic state of the ion dissociates. Figure 1d,e shows the photoelectron spectra measured in coincidence with the parent ion and C2H4+, respectively. The former reflects photoionization from the highest occupied molecular orbital (HOMO) by H7–H13, which leaves the cation intact. The photoelectron spectrum measured in coincidence with C2H4+ shows a more complex distribution, which contains contributions from the ({widetilde{A}}^{+}) and ({widetilde{B}}^{+}) ionic states. In this study, we focus on the parent-ion channel.

    Figure 2 shows the measured PECD and its coherent control by changing the XUV-IR delay with attosecond precision for the parent-ion channel. The energy (Ek)-resolved and angle (θ)-resolved PECD distribution is defined by 2(IS(Ek, θ) − IR(Ek, θ))/(IS(Ek, θ) + IR(Ek, θ)), in which IS/R(Ek, θ) is the PAD of the S/R enantiomer. The chiral nature of PECD has also been verified by switching the XUV helicity. Figure 2a,b shows the measured PECD distributions obtained from the XUV field only and the delay-averaged result in the XUV + IR two-colour fields, respectively. For the XUV-only PECD distribution, the four dipole-shaped concentric rings correspond to the four main peaks shown in Fig. 1d. After introducing an IR field (Fig. 2b), sidebands (SBs) appear between the main peaks and inherit the PECD sign of the neighbouring main peaks but with some angular modulations, a consequence of the fact that extra partial waves are involved in the two-photon-ionization process (see the ‘Theoretical methods’ section for the details). The asymmetry parameters are obtained by fitting trigonometric functions to the angle-resolved PECD, with the results shown in Extended Data Fig. 1. Figure 2c compares the extracted PECD values in the XUV-only and XUV + IR cases. In the XUV-only case, the highest PECD amounts to about 10%. In the presence of a linearly polarized IR field, the PECDs at the SB positions are enhanced by up to 5% compared with the average PECD of the neighbouring main bands. In the presence of a circularly polarized IR field, the SB PECDs are increased much more, which is in line with predictions from ref. 4 and simulations performed for MeOx. Typically, the PECD values are doubled compared with the average of their neighbouring main bands, as indicated by the black arrow in Fig. 2c. In the XUV + IR field, the PECD can also be actively controlled. Figure 2d shows the delay-resolved PECD in the presence of a linearly polarized IR field. The PECDs at all SB positions are modulated with a period of 1.33 fs, with a maximal modulation depth at SB 14, at which the PECD even changes sign. The phase shifts between the PECD oscillations of different SBs are dominated by the attochirp of the XUV APT. The delay-dependent PECDs in the case of a circularly polarized IR field is shown in Extended Data Fig. 2.

    Fig. 2: Two-colour enhancement and attosecond coherent control over PECD detected in coincidence with the parent ion.
    figure 2

    a,b, Measured angle-resolved and energy-resolved PECD distributions in the XUV-only case and the XUV + IR case with linearly polarized IR field, respectively. c, Comparison of PECD values between XUV-only photoionization and the (delay-averaged) XUV + IR two-colour photoionization, including both linear and co-rotating circular polarizations of the IR field. d, Measured PECD values at SBs 8, 10, 12 and 14 as a function of the XUV-IR delay for a linearly polarized IR field. The photoelectron signals have been integrated over a window of width 0.5 eV centred at the peaks of the main bands and SBs. The insets above panel d illustrate the PECD reversal observed in SB 14.

    We now discuss the measurement of chiral asymmetries in molecular photoionization delays, starting with the case of a linearly polarized IR field. As in the case of PECD, this approach isolates the chiral signatures of the one-photon XUV transition and eliminates possible chiral contributions from continuum–continuum transitions driven by the IR field4. Figure 3a,c shows the energy-resolved RABBIT traces obtained by integrating over the photoelectron emission angle θ from 0° to 90° (forward-emitted photoelectrons) or from 90° to 180° (backward-emitted photoelectrons). The phases of the photoelectron-yield oscillations at the SB positions have contributions from both the attochirp and the molecular photoionization delays. By evaluating the phase difference between forward-emitted and backward-emitted photoelectrons, the influence of the attochirp is cancelled, isolating the chiral asymmetry of the molecular photoionization delays. Figure 3d compares the yield oscillations of forward and backward photoelectrons for different SBs, after removal of the slowly varying background by Fourier transformation (see Methods for details). The forward photoelectrons are temporally behind the backward photoelectrons in the case of R-MeOx, that is, the SB maxima occur at larger XUV-IR delays, an effect that is most prominent for SB 8.

    Fig. 3: Chiral asymmetries in photoionization time delays.
    figure 3

    Data were acquired with a linearly polarized IR along the y direction and the photoemission angle ϕ was integrated over from −15° to +15°. a,c, RABBIT traces of backward electrons (θ is integrated over from 90° to 180°) and forward electrons (θ is integrated over from 0 to 90°), respectively, as shown in panel b. d,e, Measured and simulated energy-integrated RABBIT signals for SBs 8, 10 and 12, for which the constant background signals have been removed using Fourier transformation. The electron-energy integration width is 0.5 eV centred at the peak positions. f, Experimental and theoretical forward–backward photoionization time delay as a function of photoelectron energy for both enantiomers of MeOx. Note that the experimental results in panels a, c, d and the theoretical results in panel e correspond to R-MeOx and the data for S-MeOx are shown in Extended Data Fig. 3. a.u., arbitrary units.

    The same measurements performed on the other enantiomer (S-MeOx; Extended Data Fig. 3) showed opposite delays, demonstrating their chiral nature. Figure 3f shows the measured photoionization time delays for both enantiomers. We find that the forward–backward time delays, expressing the temporal separation of the forward-emitted and backward-emitted photoelectron wave packets, amounts to about 60 as at SB 8 and decreases with electron kinetic energy, which reflects the decreasing sensitivity of the photoelectron wave packet to the chirality of the molecular potential, as in the case of PECD amplitudes. The calculated time delays are slightly smaller than the measured ones but lie within the error bars of the measurement. The chiral nature of these forward–backward time delays is confirmed by the opposite signs of the time delays obtained for the two enantiomers in both experiment and theory.

    The molecular photoionization dynamics are simulated by adapting the theoretical framework of ref. 4, which suggested making use of RABBIT to realize the coherent control of chiral signatures in the PAD. The simulations are performed in the frozen-core static-exchange and electric-dipole approximations, following refs. 3,4,27 with the modifications described in Methods to remedy the effect of artificial resonances caused by the discretization of the photoelectron continuum and to account for the experimental conditions.

    The original theory proposal discussed photoelectron interferometry for the model chiral molecule CHBrClF (ref. 4). Compared with CHBrClF, MeOx shows a more isotropic nature, as can be seen from the smaller number of angular basis functions needed to represent the molecular wavefunctions for MeOx compared with CHBrClF (see Methods). This suggests that the photoelectrons of CHBrClF experience higher anisotropy in the molecular potential. To what extent does this influence the robustness of the control scheme? Simulations for CHBrClF in ref. 4 showed that the maximum PECD signal could be enhanced by a factor of five by optimizing the XUV-IR delay, whereas in the simulations for MeOx, optimizing the delay can increase the PECD by about a factor of two, which is a similar enhancement as the delay-averaged experimental results presented in Fig. 2. This may indicate that the presented control scheme could benefit from a less isotropic nature of the molecule. A similar argument can be made when analysing the dependence of the chiral signatures on the photoelectron energy. The PECD as well as the forward–backward time delays decrease with the SB order as shown in Figs. 2c and 3f. Nevertheless, introducing the circularly polarized IR pulse increases the PECD by about 50% for the highest SB (SB 14) compared with the PECD signal of the closest harmonic (H13) in the XUV-only case. The calculation of chiral photoionization delays has been reported in a recent publication28.

    In addition to the chiral forward–backward time delays, the 3D momentum resolution of our experiment allows us to obtain the photoionization time delays with angular resolution. Momentum-vector resolution is a prerequisite for a quantitative measurement of chiral photoionization delays because the latter depend on both laboratory-frame angles of photoemission (θ and ϕ) in any chiral-sensitive experiment. Extended Data Fig. 4 illustrates this fact by showing that the phase of the SB oscillation linearly depends on ϕ in the case of co-rotating XUV and IR fields. This is the RABBIT analogue of the angular streaking principle. Taking this property into account allows us to perform a quantitative analysis of angle-resolved chiral photoionization delays. Figure 4a,b shows the measured θ-resolved RABBIT traces of SB 8 for the two enantiomers of MeOx in the case of a co-rotating circularly polarized IR field, the configuration that yields the largest PECD enhancement in line with theoretical calculations. The RABBIT fringes are not vertical but show a tilt as a function of θ, in a direction that inverts when exchanging the enantiomers, again demonstrating the chiral nature of the effect. The time delays extracted at each θ angle (relative to θ = 90°) are shown in Fig. 4c. They vary by about 240 as over the investigated angular range. Figure 4d,e shows the calculated angle-resolved RABBIT traces for the two enantiomers, respectively, closely reproducing the opposite tilts of the RABBIT fringes for the two enantiomers. The angle-resolved photoionization delays shown in Fig. 4f closely reproduce the measured variation of the photoionization delays by about 240 as over the measured angular range. The remaining differences between the theoretical results and the experimental data may originate from non-dipole effects, non-perturbative effects in the light–molecule interaction or electronic correlation effects. The experiment was performed well within the perturbative regime and the results shown in Fig. 2a,b do not provide evidence for strong non-dipole effects. We therefore expect electronic correlations to be the primary cause for deviations between experiment and simulations.

    Fig. 4: Angle-resolved photoionization time delays in co-rotating XUV + IR fields.
    figure 4

    a,b, θ-resolved RABBIT traces of SB 8 for the R and S enantiomers of MeOx. c, The extracted photoionization time delay from a and b, for which the delay values at θ = 90° were chosen as reference. Note that the negative value of the ionization time delay indicates that the photoelectron wave packet is delayed with respect to emission at θ = 90°, chosen as the reference in these measurements. The θ-dependent maxima are highlighted in a and b as dashed lines and the equivalent lines are shown in d and e. The uncertainty of the extracted time delay (error bar) is estimated by the background-over-signal approach34. df, The corresponding results from theoretical calculations described in the text and in Methods.

    The comparison of angle-resolved time delays obtained with circularly or linearly polarized IR pulses also provides insights into the contribution of the continuum–continuum transitions to the chiral asymmetries of the time delays. Extended Data Fig. 5 shows such a direct comparison, with panels a and b displaying the raw experimental data for the case of a linearly polarized IR field and panels c and d comparing the angle-dependent photoionization time delays for linearly and circularly polarized IR fields, respectively. Whereas the delay variation amounts to approximately 180 as in the case of a linear IR field, it reaches approximately 240 as in the case of a circularly polarized IR field. To quantify the chiral asymmetry contributed by the continuum–continuum transitions, panel e shows the symmetrized time delays for a given polarization configuration (obtained as (τ(θ)S-MeOx + τ(180° − θ)R-MeOx)/2) and panel f shows their difference, which quantifies the chiral contribution of the continuum–continuum transitions to the measured photoionization delays. Although the error ranges overlap with zero, a clear trend can be recognized, corresponding to a variation of roughly 60 as over the whole angular range that we assign to the chirality of the IR-induced continuum–continuum transitions in the chiral potential.

    A comparison of the present results, obtained with circularly polarized attosecond pulses in the perturbative regime, with previous results, obtained with femtosecond pulses in the strong-field regime17,20,22, further highlights the achieved advances. The perturbative nature of the present approach results in a transparent control mechanism and good agreement with calculations. In the present work, the naturally occurring PECD effect has been enhanced by a factor of two, reaching up to 16%, and has been coherently controlled on the attosecond timescale in the perturbative regime involving pathways driven by one XUV and one IR photon each. Using femtosecond pulses, Rozen et al. observed asymmetries of up to 0.5% and their coherent control using non-perturbative two-colour strong-field ionization, but a detailed understanding of the mechanism as well as agreement with theory were missing20. In this work, we have measured chirality-sensitive photoionization delays by advancing the established RABBIT technique to circularly polarized attosecond pulses. The photoionization delays have been measured with 3D momentum resolution, in coincidence with specific ionic fragments and resolved in terms of the cationic final states. Good agreement with theory and a transparent explanation of the underlying mechanisms have been provided. By contrast, Beaulieu et al.17 have used strong-field ionization based on non-perturbative two-colour femtosecond pulses. They have measured the projection of the PAD on a 2D detector parallel to the propagation direction of the laser pulses. This geometry does not resolve the emission angle in the polarization plane of the laser pulses on which the photoionization delay depends. Moreover, the photoionization delays determined in two-colour strong-field ionization depend on the intensity of the laser pulses29, such that they do not represent intrinsic molecular properties.

    In conclusion, we have introduced attosecond chiroptical spectroscopy and applied it to resolve and control the photoionization dynamics of chiral molecules on their natural, attosecond timescale. Our study demonstrates the powerful capability of the combination of circularly polarized attosecond light pulses with the 3D-momentum-resolved electron-ion coincidence detection, which allows examination of the attosecond electron dynamics induced by photoionization in an electronic-state-resolved, angle-resolved, energy-resolved and enantiomer-resolved manner. Electron scattering in a chiral molecular potential gives rise to the asymmetric scattering amplitudes with respect to the light-propagation direction, underlying the PECD phenomenon, which has high chiral sensitivity and broad applicability. Attosecond chiroptical spectroscopy has now also revealed that one-photon-ionization delays of chiral molecules also show chiral asymmetries, defining a general approach to time resolve the electron-scattering dynamics in chiral molecules. Attosecond chiroptical spectroscopy also has the potential of answering a very important question about the origin of the chirality-induced spin selectivity (CISS) effect30, which still lacks a quantitative explanation. Whereas traditional approaches view CISS as a purely electronic (spin–orbit) effect, such models underestimate the CISS effect by typically two orders of magnitude31. Very recent models suggest that coupled electronic-nuclear dynamics could play a central role32. Attosecond chiroptical spectroscopy may offer a solution to this intriguing puzzle through its ability of temporally separating electronic from structural dynamics33.

    Continue Reading

  • Experiment sheds light on the origin of life, supporting the existence of a ‘thioester world’ before living beings – EL PAÍS English

    Experiment sheds light on the origin of life, supporting the existence of a ‘thioester world’ before living beings – EL PAÍS English

    1. Experiment sheds light on the origin of life, supporting the existence of a ‘thioester world’ before living beings  EL PAÍS English
    2. Thioester-mediated RNA aminoacylation and peptidyl-RNA synthesis in water  Nature
    3. Simple chemistry helps explain the origin of life, new study suggests  The Washington Post
    4. Scientists Make Breakthrough in Solving the Mystery of Life’s Origin  404 Media
    5. Chemists show how RNA might have started to make proteins on early Earth  Phys.org

    Continue Reading

  • Global phenology maps reveal the drivers and effects of seasonal asynchrony

    Global phenology maps reveal the drivers and effects of seasonal asynchrony

    Overview of software, data and workflow

    We conducted our LSP mapping workflow using Google Earth Engine (GEE) (v.0.1.404 or later)65 and performed additional analyses using Python66 with a set of core scientific packages (numpy67, shapely68, pandas69, geopandas70, rasterio71, xarray72, rasterstats73, dask74, scipy75, scikit-learn76, statsmodels77 and matplotlib78). All of the datasets used in our study are summarized in Supplementary Table 1, and our entire mapping workflow is summarized in Extended Data Fig. 1a.

    LSP datasets

    We used GEE to model LSP in two independent time series of remote sensing indices that are strong correlates of seasonal variability in plant productivity—NIRV, an index of the fraction of incident near-infrared light that is reflected by vegetation20, and SIF, an index of the quantity of incident photons that are absorbed by chlorophyll and re-emitted as fluorescence21. We used a 20-year time series of MODIS-derived NIRV data (daily data from 2001 to 2020, subsampled to every 4 days for computational tractability) as our main dataset (the full workflow diagram is shown in Extended Data Fig. 1a). Following best practices for estimation of patterns at the annual timescale, we chose the MCD43A4 v061 dataset34, a 16-day temporal composite79 of nadir bidirectional reflectance distribution function (BRDF)-adjusted reflectance80. We used the version of these data that is publicly available in the GEE data catalogue. We did not carry out topographic correction because the scale of our analysis (0.05°; ~5.5 km) is sufficiently coarse that spatial averaging is expected to remove topographic bias80,81. We used only pixels with quality values of ≤3 (that is, pixels for which full or magnitude-based BRDF inversions were successfully fitted82) for both the red and NIR bands (bands 1 and 2), aggregated to our target analytical resolution of 0.05° (hereafter, target resolution) using the arithmetic mean. We calculated NIRV, as described previously20, as the product of the NDVI and total NIR reflectance. NIRV values of ≤0, assumed to be invalid20, occurred predominantly in high-albedo scenes (for example, treeless snow cover; Extended Data Fig. 2d), where productivity is assumed to be minimal, so they were clamped to the minimum positive NIRV value observed during a pixel’s 20-year time series.

    To evaluate our NIRV maps, we ran some of our main analyses identically but using a global, gridded SIF dataset35. This is a roughly 4.3-year (September 2014 to January 2019), 0.05°, spatially contiguous time series dataset, interpolated by artificial neural network (ANN) from the spatially discontiguous SIF data measured along Orbiting Carbon Observatory 2 (OCO-2) orbital swaths. Rigorous internal and external validation of this dataset showed that it accurately captured the global patterns present in the original OCO-2 retrievals and that it explained 81% of the variation in contemporaneous chlorophyll fluorescence imaging spectrometer aerial measurements taken beneath OCO-2 orbits and 72% of the variation in measurements not beneath orbits35. We downloaded this dataset from the Distributed Active Archive Center for Biogeochemical Dynamics83 then ingested it into GEE.

    Given that the SIF dataset interpolates across orbital gaps but the paper describing the dataset did not explicitly validate the seasonal phenological patterns of the interpolated data, we assessed the observed seasonality in the interpolated, orbital-gap data against the observed seasonality in another, coarser-resolution SIF dataset collected by the TROPOspheric Monitoring Instrument (TROPOMI)84,85. To do so, we extracted SIF time series from the ANN-interpolated dataset at a sample of random points drawn within OCO-2 orbital gaps in three tropical realms (the Neotropics, tropical Africa, and Indo-Pacific and tropical Australia; Extended Data Fig. 6c) then compared those values to contemporaneous time series extracted from the TROPOMI SIF data. We used tropical regions for this assessment because their lack of a pronounced thermal winter creates the greatest possibility that seasonality there exhibits spatially varying patterns that are not accurately recovered by spatial interpolation from orbital-swath data. If the interpolated dataset adequately captures the true seasonal patterns of SIF within OCO-2 orbital gaps then its time series should explain the bulk of the variation in the TROPOMI time series, and it does (R2 = 0.89; Extended Data Fig. 6c).

    Data filtering

    To exclude locations where our harmonic regression-based LSP mapping methodology (see the next section) would return inaccurate results, we used an extensive filtering pipeline that removed invalid land cover, pixels with multiple types of data deficiency and pixels with statistically insignificant LSP regressions. The pixels removed from analysis by each of the filtering steps described below are mapped and summarized in Extended Data Fig. 1b.

    For land-cover filtering, we used the GEE data catalog asset for MCD12C1.06186, a MODIS product estimating annual, global land cover at our target resolution. We used the Annual International Geosphere-Biosphere Programme’s (IGBP) classification scheme (land cover type 1). To avoid low-quality data originating from non-target land cover, we excluded data from all pixels with >10% invalid land cover—including urban and built-up land, permanent snow and ice, barren land and water bodies (categories 13, 15, 16 and 0)—for all years within which that classification was assigned. Next, we retained pixels with any other land-cover classifications provided that they never switched between agricultural (categories 12 or 14) and non-agricultural (categories 1 to 11), to avoid fitting phenocycles to the noise resulting from abrupt changes between natural phenologies and those that are deliberately altered by human management (for example, irrigation). We retained pixels where land-cover assignment changed across the time series but was either always agricultural or always non-agricultural because: (1) spurious signals of change between natural land-cover types are common in regions with large, climatically driven interannual variation in plant productivity or where the actual land cover straddles categorical boundaries and challenges classification algorithms (for example, woodland, savanna and semi-arid biomes); (2) actual land-use and land-cover change (LULCC) on the ground is often too subtle to register a change in remotely sensed land-cover maps (for example, selective logging), even when it registers a clear signal in continuous metrics such as NIRV (Extended Data Fig. 2a); and (3) we only expected other forms of LULCC (for example, deforestation) to affect our modelling results in regions where different land-cover types exhibit different natural phenologies in response to the same broad bioclimatic controls, in which case pixels subject to LULCC should generate model fits that are intermediate to the phenocycles typical of the before and after land-cover types, introducing some noise into our map but neither preventing interpretation of its overarching patterns nor invalidating significant statistical results.

    While the LSP of agricultural regions is of interest in many contexts, anthropogenic LSP patterns caused by irrigation and other intensive land management practices3 could confound our phenological asynchrony analyses, which focus on the climatic drivers and evolutionary implications of longstanding, naturally occurring LSP gradients. Because of this, we used a stricter masking procedure for all datasets used to calculate LSP asynchrony maps and to run asynchrony-related analyses, omitting data from all agricultural pixels (IGBP categories 12 and 14; Extended Data Fig. 1b).

    To preclude poorly fitted LSP regressions that could cause spurious results, we removed any target-resolution pixels with data that did not satisfy a set of strict non-missingness criteria. First, we removed any pixels whose LSP time series had >50% missing data, a simple step to remove sites with data dropout because of substantial cloud contamination or MODIS quality control problems. Next, we removed any pixels without at least 10% monthly mean data availability in every month of the year. Finally, owing to a tendency for the harmonic regression procedure (described below) to interpolate spurious second LSP peaks into extended, seasonally repeating periods of missing data (for example, during high-latitude winters, when Terra and Aqua overpasses occur outside daylight hours for numerous weeks), we removed any pixels for which the binary time series of data availability (0 = missing data, 1 = data available) had a Pielou’s evenness87 of less than 0.8. We calculated Pielou’s evenness, J′ = H′/Hmax, using H′ (Shannon’s diversity index88) calculated with 12 values, each value being a monthly average proportion of non-missing 4-day-interval data over the 20-year NIRV time series. Manual inspection of fitted phenological patterns after applying this series of filtering steps confirmed successful removal of locations that would otherwise produce spurious results. These last two steps removed all locations north of roughly 60° (Extended Data Fig. 1b) because the lack of winter daylight during satellite overpass creates long, seasonally repeating stretches of unavailable data. This is also a known complication for other remote-sensing products (for example, MOD44B.061 Vegetation Continuous Fields89), but it does not affect our major findings because the same orbital physics that causes this issue also produces strong, zonally consistent temperature and photoperiod control over annual phenologies at these latitudes and, therefore, limited potential for phenological asynchrony.

    Finally, we used the harmonic regression procedure described below not only to calculate characteristic annual LSP patterns but also to estimate the significance of those patterns and filter out pixels with insignificant regression results, using a Monte Carlo framework. To do this, for each pixel in our global NIRV dataset, we randomly permuted the original LSP time-series image stack, scrambling any true seasonal signal, then ran the harmonic regression and stored an image of the R2 values at all pixels. Next, we calculated from all of the stored R2 images a single summary image of empirical P values indicating, for each pixel, the proportion of permutations for which the permuted time series’ R2 values exceeded the R2 value from the unpermuted harmonic regression. We ran this harmonic regression permutation test using 20 permutations at every pixel globally (because of computational limitations), then filtered out any pixels with an empirical P ≥ 0.05.

    Modelling of LSP

    We used harmonic regression to model the long-term average annual LSP pattern (that is, phenocycle) of every pixel in the global, filtered NIRV and SIF datasets. In our model each pixel’s full time series is predicted as a function of time as:

    $$y={beta }_{0}+{beta }_{t}t+{beta }_{1}sin ({t}_{{rm{ann}}})+{beta }_{2}cos ({t}_{{rm{ann}}})+{beta }_{3}sin ({t}_{{rm{sem}}})+{beta }_{4}cos ({t}_{{rm{sem}}})+{epsilon },$$

    where y is either the SIF or NIRV time series, t is the linear time component (days from the start of the time series), and tann and tsem are circular time expressed in annual (ann) and semiannual (sem) frequencies (that is, the day of year expressed in radians, where 2π radians corresponds to the last day of the year for tann and to the middle and last days of the year for tsem). We then retained all of the resulting coefficient maps except βt (the trend), yielding a stack of five coefficient maps that represents the detrended, long-term, characteristic annual LSP pattern at each pixel globally.

    We chose harmonic regression because it is a simple, widely used and clearly interpretable approach to time-series analysis90, and because it would enable us to characterize the long-term average annual behaviour at all terrestrial locations. Our regression formulation is algebraically equivalent to detrending the full 20-year time series, then running a Fourier transform that includes both annual and semiannual frequency components91. We designed a number of the data-filtering approaches described above to ensure against the spurious interpolation into seasonally repeating data gaps that could otherwise be caused by this method. We chose to include both the annual and semiannual frequencies in the harmonic regression to strike a balance between model complexity and overfitting. We expected that complex annual LSP patterns would occur in locations that have bimodal seasonal precipitation patterns (that is, two rainy seasons)50 and no winter freeze47. Indeed, preliminary analysis revealed numerous regions with stronger bimodal than unimodal annual LSP patterns (that is, regions containing many pixels whose R2 values were higher in semiannual-only harmonic regression models than in annual-only models). The linear combination of annual and semiannual harmonic regression components is complex enough to represent annual LSP curves that are unimodal, evenly bimodal (two equal peaks and troughs) or unevenly bimodal (featuring major and minor peaks and troughs), but not more complex, and therefore avoids overfitting by excluding unfounded higher frequencies90.

    While frequency-specific phase and amplitude estimates could be recovered from the fitted coefficients of our models, their comparative interpretation across such a wide range of phenological patterns would be difficult. Thus, for all downstream analysis and visualization, we instead use Euclidean distances and multivariate statistics calculated directly on the fitted phenocycles, which can be calculated as the multiplication of a pixel’s fitted harmonic regression coefficients with the 1-year matrix of daily time values expressed in linear time and in annual and semi-annual cyclical time. Extended Data Fig. 2a–d pairs multivariate visualization (methods described below) with demonstrations of the phenocycle-fitting procedure in various test regions, and Extended Data Fig. 2e shows a similar visualization screenshotted from the GEE app that we created for public exploration of our results (the link to which is provided within the GitHub repository for this project; https://github.com/erthward/phen_asynch, https://doi.org/10.5281/zenodo.15671259)92.

    Evaluation of LSP mapping

    We first evaluated the annual NIRV LSP map by calculating and inspecting a map of R2 values between the fitted NIRV and SIF phenocycles at all pixels (Extended Data Fig. 6b). We also checked the distribution of unimodal and bimodal phenologies against prior studies. To do this, we min–max scaled each pixel’s phenocycle to the [0, 1] interval and rotated it to start at its minimum value (to avoid problems arising from phenocycle peaks that straddle the start of the calendar year). We then extracted the heights of each phenocycle’s peaks, using the ‘find_peaks’ function in the ‘signal’ module of the Python package scipy (v.1.13.0)75, and used the absolute difference of those heights as an indicator of where a pixel lies on a spectrum between perfectly bimodal (0: indicating two peaks of equal height) and unimodal (1: assigned to phenologies having only a single peak). We mapped this index (Extended Data Fig. 6a), then visually compared it to previously published depictions of the global distribution of regions with one versus two growing seasons (see figure 3 of ref. 4).

    We also evaluated the fitted phenocycles for both LSP datasets (NIRV, SIF) by comparison with average phenocycles fitted identically to time series of PhenoCam36 NDVI and FLUXNET201537,50 GPP. For the PhenoCam analysis, we used a combination of the R93 package phenocamapi (v.0.1.5)94 and custom Python code to download all available (as of 5 March 2025) 3-day summary NDVI datasets from all cameras and regions of interest (ROIs; masked areas of uniform vegetation within a camera’s field of view, which are used to generate separate time series datasets). We used NDVI because its phenological signal, which can diverge from that of the green chromatic coordinate in some systems36, provides a better comparator to our NDVI-derived NIRV data. We used the 3-day summaries because they have reduced noise, and we analysed the 75th-percentile NDVI summary values to strike a reasonable trade-off between the tendency of higher-percentile values to be less noisy under variable lighting conditions and the risk that very high percentiles can cause outlier influence95. We dropped any camera sites that PhenoCam reports as belonging to any of the invalid IGBP land cover classes that we filtered out of our LSP analysis (urban and built-up land, permanent snow and ice, barren land and water bodies) or as being agricultural (because agricultural management could cause an entirely different phenology within a camera’s field of view than the spatially averaged phenological signal reflected in our LSP map), leaving a total of 368 camera sites eligible for analysis. Before fitting a harmonic regression for each site, we removed outliers from each of the site’s ROI datasets (using the outlier flag provided by PhenoCam), then combined all datasets by averaging each day’s values across all ROIs to approximate the integrated land cover signal in our LSP dataset at that site. We then used the same harmonic regression model used in the LSP-fitting procedure described above to calculate a set of five coefficients describing the detrended, average annual NDVI phenology for a site, and from those coefficients performed matrix multiplication to recover the fitted characteristic annual NDVI phenocycle for each site. Finally, for each site, we calculated the R2 values between the site’s characteristic annual NDVI phenocycle and the LSP phenocycle corresponding to the site (that is, the pixel where the camera is located or, if that pixel is masked in our LSP dataset, the nearest valid pixel within a two-pixel-wide box that surrounds it). We summarized this evaluation procedure across all camera sites by producing, for each LSP dataset: (1) a scatter plot of the LSP-NDVI R2 values plotted on the Whittaker biomes96, to depict bioclimatic patterns in evaluation performance; and (2) a scatterplot comparing LSP-NDVI R2 values to NDVI time series lengths, to depict the relationship between camera data availability and evaluation performance (Extended Data Fig. 6d).

    For the FLUXNET2015 comparison, we manually downloaded all datasets available at the time of access (11 October 2021), then, as with PhenoCam, dropped all flux tower sites reporting invalid and agricultural land cover types, yielding 170 valid GPP datasets for analysis. Before fitting a harmonic regression to each dataset, we first removed all datapoints with a daily quality value of <0.7 (that is, with <70% measured or good-quality gap-filled data contributing to their daily aggregated values). We then used the same methods as described for the PhenoCam NDVI comparison above to fit a harmonic regression, predict a characteristic annual time series, calculate R2 values between the annual time series and those from their closest available LSP pixels (up to 2 pixels distant, otherwise a tower’s dataset was dropped) and visualize the results (Extended Data Fig. 6e).

    LSP visualization

    To visualize the global variability of seasonal LSP that is present in the results of our harmonic regression, we used colour-composite visualization of the results of a dimensionality-reduction analysis to produce a single global map. First, we used Python v.3.7 and the eofs package (v.1.4.0)97 to run EOF analysis on the covariance matrix of the global set of NIRV phenocycles. We standardized each pixel’s phenocycle before EOF calculation, ensuring that all pixels had equal variances of 1 and therefore allowing the EOF analysis to highlight global variability in the shape and timing of LSP patterns, our topic of interest, irrespective of spatial variation in NIRV amplitude. Following common practice in EOF analysis, we used the square root of the cosine of the latitude as pixel area weights.

    This calculation reduced the global diversity of average annual LSP patterns to four EOFs. Finding that the first three EOFs cumulatively explain >90% of the variation in the dataset (91.62%; Extended Data Fig. 4a), we min–max scaled them, then displayed them using the RGB colour channels, visualizing the bulk majority of global LSP variability within a single map. As they have embedded within them both the unremarkable north–south hemispheric seasonality dipole and hemisphere-independent patterns of interest (for example, monsoon-driven LSP dynamics), we transformed the raw EOF maps before RGB visualization to represent phenological variability in a globally consistent colour scheme. To accomplish this, we used WebPlotDigitizer98 to digitize a geospatial vector file of the mean ITCZ in both boreal summer (June, July, August) and boreal winter (December, January, February)38, then calculated a single, annual mean ITCZ vector by averaging the boreal summer and winter latitudes at evenly spaced longitudes around the globe. Finally, for each EOF, we constructed a synthetic, transformed map by calculating w × EOF + (1 − w) × (1 − EOF), where w varies from 1 in the northern hemisphere to 0 in the southern hemisphere and transitions linearly from 1 to 0 within a 10° latitudinal band surrounding the annual mean ITCZ. We chose to use the ITCZ as the latitudinal boundary across which to transform the EOF maps because it serves as a more natural meteorological Equator than does the geographical Equator17,38. To help to interpret the result of this visualization across the region surrounding the ITCZ (Fig. 1), where some colour-warping occurs, we also generated RGB composite maps using untransformed EOF maps and using EOF maps transformed uniformly as 1  EOF (Extended Data Fig. 4b,c). As this transformation is used only for visual comparison across hemispheres, it has no influence on any of the analytical results reported in our work.

    To depict the characteristic phenocycles corresponding to the RGB visualization, we use mini-batch k-means clustering (a version of the standard k-means clustering algorithm that reduces computational burden by using only a fixed-size random subsample of the full dataset at each iteration) to cluster the standardized, fitted phenocycles within a region into k colours, for k = 1:12, then visually inspect a scree plot to determine the optimal value of k. Using that chosen value, we assign each pixel to one of k clusters, then plot each cluster centre (after min–max scaling) as its characteristic phenocycle, coloured by the median RGB value across all pixels in the cluster. We used this procedure to produce plots interpreting the predominant phenocycles both globally (Fig. 1 (main)) and within various focal regions (Fig. 1a–d and Extended Data Fig. 5). Before clustering the global map, we rotated the fitted phenocycles of all pixels below the mean ITCZ by 182 days (that is, half a year) to allow similar phenologies in the northern and southern hemispheres to cluster together.

    Discovering regional phenological variability in the Great Basin of the United States that appeared to match the cheatgrass-invaded, sagebrush and montane phenologies presented previously42, we used ancillary data from ref. 43, aggregated to the target resolution of our map, to calculate the average estimated percentage of annual herbaceous cover in each of the three predominant clusters depicted in our analysis (Fig. 1b). To support our interpretation of the three clusters as annual-invaded communities, sagebrush and montane vegetation, which we based on the differences in their estimated average annual herbaceous cover and on a visual comparison to ref. 42, we used ANOVA to test for a significance difference in the estimated percentage of annual herbaceous cover across all three clusters, followed by a Tukey’s honest significant difference test to test for significant pairwise differences between the clusters.

    To better highlight complex geographical patterns of spatially variable LSP timing, we also produced a video (Supplementary Video 1) animating the min–max scaled average NIRV phenocycle at each pixel. Scaling each pixel’s phenocycle in this way forces all pixels to a common annual amplitude (from zero to one), ignoring spatial differences in intra-annual variability caused by variable ecosystem productivity, and thus highlighting spatial differences in the timing and rates of change of LSP.

    Calculation of phenological asynchrony

    We exported the GEE results of our filtered harmonic regression as a global set of tiled, multiband images of regression coefficients. We used GEE’s TensorFlow output format and ‘kernelSize’ argument to generate tiles that overlapped their neighbours by 300 km (double the largest neighbourhood size in our asynchrony calculations), to allow asynchrony to be calculated independently and in parallel.

    For each LSP dataset (NIRV, SIF), we calculated our asynchrony metric pixel-wise, for all pixels with at least 30 available neighbours, using an algorithm based on Martin et al.12 and depicted in Extended Data Fig. 7a:

    1. (1)

      Calculate the standardized phenocycle for a focal pixel.

    2. (2)

      Identify all pixels of which the centrepoints are within the chosen neighbourhood radius of the focal pixel (the neighbour pixels).

    3. (3)

      For each neighbour pixel: (a) calculate its standardized phenocycle; (b) calculate the 365-dimensional Euclidean phenological distance between its phenocycle and the focal pixel’s phenocycle; (c) calculate its geographical (geodesic) distance to the focal pixel.

    4. (4)

      Calculate asynchrony as the slope of the regression of Euclidean phenological neighbour distances on geographical neighbour distances (or as zero, wherever the slope has P > 0.01).

    We used a regression approach to calculate the asynchrony metric because it explicitly estimates the spatial rate of change in phenology, and therefore well represents the spatial rate of change of seasonal timing that is the subject of the ASH12. We standardized phenocycles, nullifying differences in amplitude, before calculating Euclidean distances between them, therefore preserving the timing differences that we are interested in, even between similar-shape but out-of-phase curves (a criterion not met by other common distance metrics, such as dynamic time warping). We ran this calculation in Julia (v.1.4.1)99 on UC Berkeley’s Savio cluster, parallelized by tile, then mosaicked the results into a global map (Fig. 2a).

    We produced this global map for each of three neighbourhood radii (50, 100, and 150 km), enabling us to check the sensitivity of our maps and our downstream results to this decision. The values of the resulting maps, expressed as a spatial rate of change in the target variable’s units (that is, Δunittarget_variablem), scale arbitrarily with a map’s neighbourhood radius, but each map provides an internally valid quantitative basis for assessing and comparing asynchrony between sites. To assess the overall level of agreement between the NIRV and SIF asynchrony maps, despite the fine-scale noise expected in a neighbourhood metric, we mapped and scatter plotted pixel-wise comparisons between the two datasets for each of the three neighbourhood radii (Extended Data Fig. 8). Moreover, to evaluate the scale-sensitivity of the LSP asynchrony maps (and of the asynchrony maps that we likewise calculated for the climatic covariates described below), we assessed, for each mapped variable, the R2 values for all three pairwise interneighbourhood map comparisons (Extended Data Fig. 7b).

    To visually depict the asynchrony algorithm, we first simulated harmonic-regression output for a low-asynchrony region as a five-layer stack of coefficient values with rasters of low relative-magnitude Gaussian noise added to them and for a high-asynchrony region as a five-layer stack of mean coefficient values with large relative-magnitude, spatially autocorrelated noise added to them using neutral landscape models generated using the nlmpy Python package100. We represented each five-layer simulated map as a single-layer map by first calculating each pixel’s phenocycle from its simulated vector of harmonic regression coefficients, then calculating the day of the year when its simulated phenocycle attains its peak value. We used this summary map, all pixels’ simulated phenocycles and the phenological-distance–geographical-distance regression (the slope of which serves as the asynchrony metric) to graphically depict the asynchrony calculation procedure (Extended Data Fig. 7a).

    Phenological asynchrony model covariates

    For the random-forest (RF) model exploring the potential drivers of phenological asynchrony (see below), we produced rasters of physiographic and environmental covariates using workflows combining GEE, Julia, Python and GDAL (v.2.2.3)101. First, we applied the same harmonic regression and asynchrony-mapping pipeline described above, skipping the masking steps that were specific to LSP data quality concerns, to the 64-year TerraClimate time series dataset102 in the GEE catalogue, generating asynchrony maps for the climatic factors potentially driving phenological asynchrony: monthly minimum and maximum temperature, monthly precipitation and monthly climate water deficit. We supplemented this with an equivalently produced map of asynchrony in cloud cover, using cloud cover fractions calculated from the internal cloud algorithm flag bit (bit 10 of the 1 km reflectance data QA band) of the MODIS Aqua and Terra daily 1 km global surface reflectance datasets (MYD09GA.061103 and MOD09GA.06104) in the GEE catalogue. The R2 values from these harmonic regressions are also mapped in Extended Data Fig. 3, and the asynchrony of the climatic factors is mapped in Extended Data Fig. 7.

    To model the potential importance of topographic complexity for driving phenological asynchrony, we downloaded a global map of the vector ruggedness metric105. We chose this over other measures of topographic complexity because of its reduced correlation with slope. We downloaded data published previously106, choosing a map based on Global Multi-resolution Terrain Elevation Data 2010 (GMTED2010) elevation data107 and median-aggregated at a scale on par with the neighbourhood size of our main LSP asynchrony dataset (100 km).

    To allow the model to reflect phenological asynchrony between structurally distinct vegetation communities, we used GEE to create a global map of entropy in vegetation structure within 100 km neighbourhoods (hereafter, the vegetation entropy map). To do this, we used the same 20-year time series of annual MODIS IGBP 0.05° land cover86 that we used in our LSP data-filtering workflow. We reclassed land cover into categories of forest (IGBP classes 1–5: evergreen or deciduous broadleaf or needleleaf forests and mixed forest), shrubland (IGBP classes 6 and 7: closed and open shrublands), savanna (IGBP classes 8 and 9: woody savannas and savannas), grassland (IGBP class 10) or permanent wetland (IGBP class 11). We then applied the same mask used to calculate LSP asynchrony, so that the information captured by this covariate would reflect the information included in the LSP asynchrony response variable. Next, we reduced the 20-year time series to a single map representing the modal class for each pixel across all years. Finally, we produced the covariate map by calculating the entropy of the vegetation structure classes within each pixel’s 100 km radius (the neighbourhood size of our main analysis) as −ΣiP(ci) log2P(ci), where c is vegetation structure class and P(ci) is the proportion of the neighbourhood that is assigned class i.

    As LSP asynchrony patterns could be influenced by human LULCC, we used GEE to create two other 100-km neighbourhood covariates: the mean proportion of subpixels classified as LULCC, and the mean frequency of fire. We derived the mean LULCC proportion map from a global, harmonized map of Landsat-resolution (30 m) land-cover change and land use in 2019108. In GEE, we calculated the proportion of subpixels within each of our target-resolution pixels that were classified as any land-cover change or land-use class, including classes 92–116 and 212–236 (tree cover loss since 2000, with or without regrowth), classes 240–249 (built-up land) and class 252 (cropland). We applied to that LULCC proportion map the same mask used to calculate LSP asynchrony, then calculated the mean within a 100 km radial neighbourhood for every pixel.

    As the source data for the LULCC map explicitly excludes fire-driven tree cover loss, we also used GEE to produce a separate covariate map estimating the neighbourhood mean frequency of fire. To do this, for each pixel we counted the number of months with a recorded burn date in the global monthly MODIS Burned Area dataset, MCD64A1.061109, divided that by the total number of months in the dataset, used the arithmetic mean to aggregate the map from its original 500 m resolution to our target resolution, applied the same mask applied to the map used to calculate LSP asynchrony, then calculated the mean of that fire frequency map within each pixel’s 100 km radial neighbourhood.

    Modelling phenological asynchrony drivers

    To explore the potential drivers of LSP asynchrony, we constructed an RF model using the R package ranger (v.0.13.1)110 and incorporating the set of covariates described above, formulated as:

    $${{rm{LSP}}.{rm{asy}}}_{{rm{neigh}}} sim {{rm{ppt}}.{rm{asy}}}_{{rm{neigh}}}+{{rm{tmp}}.min .{rm{asy}}}_{{rm{neigh}}}+{{rm{tmp}}.max .{rm{asy}}}_{{rm{neigh}}}+{{rm{def}}.{rm{asy}}}_{{rm{neigh}}}+{{rm{cld}}.{rm{asy}}}_{{rm{neigh}}}+{rm{vrm}}.{rm{med}}+{rm{veg}}.{rm{ent}}+{rm{luc}}.{rm{prp}}.{rm{mea}}+{rm{brn}}.{rm{frq}}.{rm{mea}}[+x+y],$$

    where LSP.asy is asynchrony of the LSP dataset used in a given model (either NIRV or SIF), ppt.asy is PA, tmp.min.asy and tmp.max.asy are minimum and maximum temperature asynchrony, def.asy is climate water deficit asynchrony, cld.asy is cloud cover asynchrony, neigh indicates the asynchrony neighbourhood radius used to calculate the asynchrony metrics for a given model (50, 100 or 150 km), veg.ent is vegetation structural entropy, vrm.med is the median vector ruggedness metric, luc.prp.mea is mean proportion of LULCC, brn.frq.mea is mean fire frequency, and x and y are pixel longitude and latitude (within brackets to indicate their inclusion in only half of the suite of models). We chose the RF algorithm owing to its ability to robustly model nonlinear relationships, suited to our expectation that phenological asynchrony would be driven by different and potentially interacting factors in different regions of the globe. We developed a comprehensive and conservative modelling workflow, which we ran once for each combination of LSP dataset (NIRV, SIF), neighbourhood radius (50 km, 100 km, 150 km), and coordinate inclusion (geographical coordinates either included or excluded as covariates). We examined the sensitivity of our RF models to the inclusion of geographical coordinates because of the lack of consensus about how to handle spatial data in RF modelling111,112. This produced a final set of 12 models (Extended Data Fig. 9a). As we found that salient results were largely insensitive to choice of LSP dataset, neighbourhood radius and coordinate inclusion, we chose the 100 km, NIRV-based, coordinates-included model as the main model to summarize and discuss in the main text of this article.

    Before producing final results, we used R v.4.0.3 to prepare the modelling data, tune hyperparameters and carry out feature selection. First, we projected the response and covariate rasters to a metric projection (EPSG:3857) to ensure that coordinates were expressed in metres, then stacked them and extracted their values at all valid (that is, non-masked) pixels. Next, we carried out comprehensive hyperparameter tuning113, assessing model performance as a function of five RF tuning parameters (number of trees per forest: ‘ntree’ = 150, 200, 250, 300; fraction of observations to use in each tree, for tree decorrelation: ‘sample.fraction’ = 0.3, 0.55, 0.8; minimum number of observations that can be captured by a node: ‘min.node.size’ = 1, 3, 5, 10; size of random subset of variables from which to choose each node’s split variable: ‘mtry’ = 1, 3, 5; and whether to sample with replacement: ‘replace’ = true, false) and as a function of the fraction of the full global dataset used for modelling (‘subset.frac’ = 0.05, 0.005; drawn as a random subsample, quartile-stratified by the LSP response variable, to reduce the computational demand imposed by the size of the modelling dataset without causing excessive information loss). We included geographical coordinates in all models used for hyperparameter tuning, as we intended to retain them in the main model unless we found that predominant results were highly sensitive to their inclusion. We used as a performance metric the root mean squared error (r.m.s.e.) of the model fitted to a 60% training split of the subsampled global dataset and found that the r.m.s.e. of the predictions made on the 40% test split yielded the same set of optimum-performance hyperparameter choices. Lastly, before running the final set of models, we confirmed that none of our subsetted datasets contained variables with a collinearity of R2 ≥ 0.75, and we used the Boruta feature-selection algorithm and R package (Boruta v.7.0.0)114 to select our final feature set (but found no features that should be dropped).

    We constructed the final 12 models using the optimum hyperparameters indicated by our tuning results (ntree = 300, sample.fraction = 0.8, min.node.size = 1, mtry = 5, replace = false and subset.frac = 0.05). To evaluate each model, we calculated two variable importance metrics—ranger’s default permutation-based importance metric, which compares the cross-tree average accuracy of out-of-bag sample predictions to the accuracy after permuting covariate values, and the absolute SHAP values115 summed across all predictions in a model’s training dataset, calculated using the R fastshap package (v.0.0.7)116—as well as two metrics of overall model performance, R2 and r.m.s.e. To help with spatial model assessment, we used trained models to make LSP asynchrony predictions at all global pixels, then calculated prediction error maps (Extended Data Fig. 9b shows the error map for the main model). Lastly, to aid spatial interpretability of the models, we calculated pixel-wise SHAP values and produced global SHAP maps for each covariate.

    Noting low variability across models in the covariates identified as having the highest importance (Extended Data Fig. 9a), we summarized the main model (100-km NIRV asynchrony, coordinates included) in the text and estimated the predominance of the top two covariates in that model, PA (ppt.asy) and MTA (tmp.min.asy), as a normalized difference of absolute SHAP values: predom = (|SHAPppt.asy| − |SHAPtmp.min.asy|)/(|SHAPppt.asy| + |SHAPtmp.min.asy|). We plotted a summary map of the normalized difference across global regions of high LSP asynchrony (that is, pixels ≥85th percentile), to show regional variation in the predominance or codominance of these two drivers (Fig. 2b; Extended Data Fig. 9c shows predominance across all covariates except geographical coordinates).

    Isoclimatic phenological asynchrony

    To test the hypothesis that phenological asynchrony is less dependent on climatic difference at low latitudes than at higher latitudes, we performed an ensemble analysis. Each sub-analysis in the ensemble first uses clustering to delineate a global set of high-asynchrony regions, then uses matrix regressions to estimate the slope of the relationship between climatic and phenological distance (hereafter, the climate–phenology correlation) within each of those regions. We defined the sub-analyses within the ensemble using unique combinations of low, middle and high values for three hyperparameters to which our final results could exhibit sensitivity, then used Monte Carlo analysis to assess the relationship, across the ensemble, between regions’ mean latitudes and the strengths of their climate–phenology correlations.

    To delineate high-asynchrony regions, we first converted our NIRV LSP asynchrony map into a map of maximum asynchrony pixels by setting all pixels ≥95th percentile asynchrony value to 1 and masking everything else. We then used the density-based spatial clustering of applications with noise (DBSCAN) algorithm117, implemented in the Python package sklearn (v.1.0.2)76, to cluster those high-asynchrony pixels. We chose the DBSCAN algorithm owing to its ability to robustly identify clusters of arbitrary shape around the high-density centres of a point set without forcing all points to have cluster assignments, which was a good match for the noisiness of our asynchrony map. Finally, we used the alpha-complex algorithm (a straight-line edge variant of the alpha-hull algorithm), implemented in Python by the Alpha Shape Toolbox (alphashape, v.1.3.1)118, to delineate high-asynchrony regions around those clusters. This enabled us to relax the convexity and contiguity assumptions of other hull-determination algorithms and, therefore, to flexibly delineate regions with complex shapes (for example, mountain arcs) without inevitably including all intervening geographical areas, as would occur with convex hulls.

    To assess the relationship between the mean latitude of a region and the strength of the climate–phenology correlation within that region, we first standardized and stacked each of the 19 WorldClim bioclimatic variables119 and standardized our global map of fitted phenocycles. Then, for each delineated region, we executed the following steps:

    1. (1)

      Draw a set of 1,000 random points within the region that all fall within non-masked NIRv LSP pixels (or draw the maximum number of points possible, if regions are too small for 1,000 points).

    2. (2)

      Calculate the matrix of pairwise phenological distances (distphen) between all points (as 365-dimensional pairwise Euclidean distances between phenocycles).

    3. (3)

      Calculate the matrix of pairwise climatic distances (distclim) between all points (as 19-dimensional pairwise Euclidean distances between bioclimatic values).

    4. (4)

      Calculate the matrix of pairwise geographical distances (distgeog) between all points (as geodesic distances).

    5. (5)

      Standardize all three pairwise distance matrix variables (so that coefficients of all regressions are β coefficients and, therefore, comparable), then run MMRR53 using the formula phenology ~ βc climate + βg geography, where βc and βg indicate the strengths of the relationships between climatic and phenological distances and between geographical and phenological distances, respectively.

    To hedge against hyperparameter sensitivity, we chose reasonable ranges of low, middle and high values of the key parameters in the clustering and hull-delineation algorithms from which to compose our ensemble. The DBSCAN clustering algorithm relies on two parameters to which our results might be sensitive: ‘eps’ (epsilon), the maximum geographical distance between two points that can be considered to be in the same neighbourhood; and ‘min_samples’, the minimum number of samples required within a neighbourhood for a point to be considered as a core point. The alpha-complex algorithm has an additional parameter to which our results might be sensitive: ‘alpha’, a value controlling how edge members are chosen and, therefore, determining the maximum complexity of a hull’s edge. To create the ensemble, we reran the full regionalization and climate–phenology correlation analysis once for each combination of the following parameter values: eps = 2, 3.5, 5; min_samples = 0.3, 0.45, 0.6; and alpha = 0.25, 0.75, 1.25.

    As a final step, we summarized the ensemble results across the 27 parameterizations by running the ordinary least squares regression model ({beta }_{{rm{c}}} sim {gamma }_{{rm{lat}}}| overline{{rm{lat}}}| ), using γlat to quantify the relationship between the absolute value of the mean latitude of each cluster and the strength of its climate–phenology correlation (Fig. 3b). As this regression violates the assumption that samples of the independent variables are IID—each point represents a clustered and delineated high-asynchrony region, and those regions can overlap across distinct parameterizations of the sub-analyses—we used Monte Carlo analysis to generate an empirical P value for γlat in the ensemble linear regression model. We ran 1,000 iterations of the same regression, each time permuting the vector of (| overline{{rm{lat}}}| ) values, then calculated an empirical P value as the fraction of the 1,000 simulated γlat that are at least as extreme as the observed γlat (Fig. 3c). To provide a spatially explicit geographical interpretation of the results of this analysis, we mapped a summary of the ensemble results as a hexbin map (Fig. 3a), with the colour of each hexbin indicating the mean βc of all high-asynchrony regions (that is, delineated alpha hulls) overlapping the bin’s hexagon.

    Allochrony by allopatry: flowering

    To explore the ability of remotely sensed LSP to predict geographical variation in flowering phenology, we tested the correlation between NIRV phenocycles and dates of flowering observations for all available iNaturalist taxa with non-unimodal flowering histograms and without extremely broad latitudinal distributions. First, we used the Python API client pyinaturalist (v.0.19.0)120 to download from iNaturalist the weekly flowering-observation histogram, and the first ≤5,000 native, non-captive, research-grade flowering observations corresponding to that histogram, for every taxon having ≥50 annotated flowering observation records at the time of download (downloads completed between 5 June 2024, 23:00 UTC and 9 June 2024, 00:00 UTC). This included a total of 7,251 taxa out of the 34,438 iNaturalist taxa with at least one observation (21.1%). We truncated the raw observation datasets to ≤5,000 per taxon to limit strain on the iNaturalist API; preliminary results showed that this decision was inconsequential because none of the 39 taxa affected would ultimately be retained for later analyses.

    We further filtered the observation points for each taxon to only those with at least 1 km positional accuracy, then used the alpha-complex algorithm118, with alpha set to 0.75 (the middle value used in our isoclimatic phenological asynchrony analysis) to fit a conservative geographical boundary (hereafter, observation range) to the set of iNaturalist observations. One taxon dropped out of our analysis at this stage because of the failure to fit an observation range. We then estimated the number of peaks in the flowering-week histogram for each taxon using the following steps:

    1. (1)

      ‘Rotate’ the histogram so that the first instance of its minimum value moved into the first position in the vector, to avoid spurious results arising from flowering peaks that straddle the last and first weeks of the calendar year.

    2. (2)

      Fit a kernel density estimation (KDE) to the histogram, using a bandwidth of 5 weeks, to reduce the noise resulting from temporal variance in observation counts.

    3. (3)

      Use a simple, neighbour-comparison-based peak-search algorithm (implemented in the find_peaks function in the signal module of the Python package scipy v.1.13.0)75 to count the number of peaks in the KDE with a height ≥60% of the overall range of values in the histogram.

    4. (4)

      Calculate the absolute value of the lag-1 temporal autocorrelation in the observed KDE and in KDEs fitted to 100 permuted versions of the rotated flowering histogram.

    5. (5)

      If the non-permuted KDE has an empirical P ≤ 0.05 (that is, if the absolute value of the lag-1 temporal autocorrelation of the non-permuted KDE is greater than that of ≥95% of the permuted KDEs), then it has a significant signal of temporal autocorrelation that probably represents non-random seasonal variability in flowering activity, so assign the counted number of peaks as the observed number of flowering-time peaks for the taxon; otherwise, assign zero as the observed number of statistically significant flowering-time peaks.

    Executing this procedure for all available taxa resulted in 6391 taxa (88.2%) with unimodal flowering-time histograms and 859 non-unimodal taxa, including 123 taxa (1.7%) with bimodal histograms, one taxon with a trimodal histogram and 735 taxa (10.1%) with no statistically significant flowering-time peaks. We dropped the unimodal taxa from further analysis because they were unlikely to exhibit the sharp geographical discontinuities in flowering phenology that were our main interest. We retained the 859 non-unimodal taxa to test for significant signals of allochrony by allopatry. We summarized these results by creating a set of hexbins covering all fitted observation ranges and then mapping, for each hexagon, the proportions of taxa with zero and with ≥2 flowering-time peaks and the overall proportion of all non-unimodal taxa (Extended Data Fig. 10). To preclude significant but uninteresting results for taxa broadly distributed across latitudes, and therefore affected by the opposite seasonalities of the northern and southern hemispheres, we dropped any taxa with samples extending beyond both 10° north and south latitudes (196 taxa).

    We then looked for evidence of allochrony by allopatry by testing each of the 663 remaining taxa for a correlation between intersite flowering-date distances and intersite LSP distances. To do this, we fitted an MMRR model for each taxon, specified as flowering_date ~ βLSP LSP + βC climate + βG geography, where the variables are pairwise distance matrices and βLSP and its P value were our output values of interest, indicating the strength and statistical significance of the relationship between LSP and flowering date distances after accounting for environmental and geographical distances. Some non-unimodal taxa may flower opportunistically, perennially or at multiple discrete times of year within the same sites, and should therefore yield insignificant βLSP values, but taxa exhibiting the strong geographical discontinuities in flowering time that we would expect under allochrony by allopatry should yield a significant, positive βLSP value. To produce the distance covariates for this model, we calculated flowering date distances as the shorter of the two forward-time or backward-time distances between two observations’ numerical day-of-year values, LSP distances as the 365-dimensional Euclidean distances between the observation sites’ NIRV phenocycles, climate distances as the 19-dimensional Euclidean distances between the sites’ vectors of standardized WorldClim119 bioclimatic variables and geographical distances as the geodesic distances between sites. We corrected βLSP P values to control for the false-discovery rate (FDR) using the ‘false_discovery_control’ function in the ‘stats’ module of the Python package scipy (v.1.13.0)75 with the Benjamini–Hochberg method. Supplementary Table 4 provides results for the 43 taxa that remained significant after FDR control (of 614 taxa successfully tested, after 49 dropped out because of insufficient data for model-fitting), and the full results from all stages of this analysis are archived with the data for this study.

    To visualize the results of this analysis for an example taxon, we plotted a temporal comparison between the flowering observation dates and the flowering observation locations’ min-max scaled phenocycles as well as a map of the observation locations, coloured according to k-means clustering of the phenocycles (k = 2) to highlight the spatial and temporal structure of the geographical discontinuity in phenology. We constructed this visualization (Fig. 4a) for two example taxa with FDR-corrected significance, chosen to demonstrate the correspondence of their patterns of allochrony by allopatry to the regional LSP patterns we had mapped and highlighted earlier in the article (M. scabra, in southwestern North America; S. parviflorum, in South Africa).

    Allochrony by allopatry: genetics

    To test whether remotely sensed LSP predicts the phenologically driven isolation by time22 that is expected to result from allochrony by allopatry, above and beyond isolation by distance121 and isolation by environment122, we fitted genetic MMRR models to a pair of datasets from two of the few published genetic studies of the ASH, substituting LSP distances calculated from our dataset for the authors’ previously used measures of asynchronous seasonality, then compared our results to theirs. First, we gathered and prepared the genomic and geographical data from the only genomic test of the ASH of which we are aware, a study of the eastern Brazilian toad R. granulosa25. We used the R package adegenet (v.2.1.5)123,124 and data downloaded from the Dryad repository for that study (https://datadryad.org/stash/dataset/doi:10.5061/dryad.pc866t1p4) to calculate a pairwise genetic distance matrix for 80 samples collected from 51 localities, based on the Euclidean distance between allele frequencies at 7,674 independent single-nucleotide polymorphism loci. We calculated geographical- and LSP-distance matrices as described above, using the geographical coordinates of each sample, and prepared a climatic distance matrix using the Euclidean distances between standardized versions of the four WorldClim119 bioclimatic variables used in the original study: annual mean temperature (BIO1), temperature seasonality (BIO4), annual precipitation (BIO12) and precipitation seasonality (BIO15). Five samples fell within masked pixels in our LSP dataset and thus could not be included in our analysis, yielding a final sample size of 75. We fit an MMRR model specified as genetic ~βLSP LSP + βC climate + βG geography, then compared our results to the results presented in table 4 of ref. 25. To visualize our findings we used k-means clustering with Euclidean distances to divide the samples into k = 2 clusters, first clustering by NIRV phenocycles, then a second time clustering by genetic distance vectors. We then prepared side-by-side equivalent plots showing sample localities and their min–max-scaled phenocycles, coloured by either of those clusterings, providing a simple visual indication of the extent to which our LSP map recapitulates the observed genetic structure (Fig. 4b (top)).

    To explore whether disparately related, sympatric taxa might exhibit similar patterns of isolation by asynchrony, we repeated the same procedure for the only other sympatric genetic dataset that we could find within previous studies of the ASH: cytochrome B sequencing data for the lesser woodcreeper (X. fuscus; Furnariidae)32. We first downloaded sample location data from the Zenodo archive for the study (http://zenodo.org/records/5012226)125 and the FASTA-formatted sample sequence data from GenBank. We aligned sequences using ClustalW (v.2.1)126 with the default parameter settings and then used jModelTest2 (v.2.1.10)127 to compare the fit of 44 models of sequence evolution to the sequence data. We then calculated pairwise genetic distances under the best-fit model (TVM + G), identified with AICc scores, using MEGA X (v.10.1.7)128. We then followed the same steps as for the R. granulosa data, except that we used all 19 WorldClim variables. Our sample size was reduced to 31 because three sampling sites fell within masked LSP pixels. The results are visualized in Fig. 4b (bottom).

    Allochrony by allopatry: coffee harvest

    To test for significant agreement between the harvest season map produced by the National Federation of Coffee Growers of Colombia (Federación Nacional de Cafeteros de Colombia, or Fedecafé) and our LSP map, we constructed a permutation-based test of an index of similarity between the harvest categories in the Fedecafé map and the categories resulting from clustering on NIRV phenocycles. First, we used WebPlotDigitizer98 to digitize and save a set of sampling points within each of the four harvest season colours displayed in a previously published Fedecafé map54. Next, we used Python to extract NIRV phenocycles at all unmasked pixels coinciding with those points and then used k-means clustering to cluster all extracted phenocycles into four clusters. We then calculated the Jaccard index129 of this cluster assignment vis-a-vis the Fedecafé harvest season assignment as:

    $$J={n}_{{rm{both}}}/({n}_{{rm{Fedecaf}}{rm{ acute{{rm{e}}} }}}+{n}_{{rm{LSP}}}+{n}_{{rm{both}}}),$$

    where nFedecafé is a count of pairwise point comparisons that have the same assignment only within the Fedecafé map, nLSP is a count of those that have the same assignment only within the LSP clustering and nboth is a count of those that have the same assignment in both datasets. Finally, we executed the same operation 1,000 times, each time first permuting the relationship between the sampling points and their phenocycles, generating a set of null J values against which to calculate an empirical P value for the observed J value (as the fraction of the 1,000 simulated J values that are at least as large as the observed J). We visualize the overall agreement between the Fedecafé map and ours by plotting sampling points on top of the RGB composite from the LSP EOF analysis (but not transformed across the ITCZ, given the colour-warping this causes within this region) and using colour to match the harvest season assignments of the sampling points to a series of line plots of their median, 10th percentile and 90th percentile phenocycles in our dataset (Fig. 4c).

    Reporting summary

    Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.

    Continue Reading

  • NASA Seeks Volunteers to Track Artemis II Mission

    NASA Seeks Volunteers to Track Artemis II Mission

    NASA seeks volunteers to passively track the Artemis II Orion spacecraft as the crewed mission travels to the Moon and back to Earth.

    The Artemis II test flight, a launch of the agency’s SLS (Space Launch System) rocket and Orion spacecraft, will send NASA astronauts Reid Wiseman, Victor Glover, and Christina Koch, along with CSA (Canadian Space Agency) astronaut Jeremy Hansen, on an approximately 10-day mission around the Moon.

    The mission, targeted for no later than April 2026, will rely on NASA’s Near Space Network and Deep Space Network for primary communications and tracking support throughout its launch, orbit, and reentry. However, with a growing focus on commercialization, NASA wants to further understand industry’s tracking capabilities.  

    This collaboration opportunity builds upon a previous request released by NASA’s SCaN (Space Communication and Navigation) Program during the Artemis I mission, where ten volunteers successfully tracked the uncrewed Orion spacecraft in 2022 on its journey thousands of miles beyond the Moon and back.

    During the Artemis I mission, participants – ranging from international space agencies, academic institutions, commercial companies, nonprofits, and private citizens – attempted to receive Orion’s signal and use their respective ground antennas to track and measure changes in the radio waves transmitted by Orion.

    Kevin Coggins

    Deputy Associate Administrator for SCaN

    “By offering this opportunity to the broader aerospace community, we can identify available tracking capabilities outside the government,” said Kevin Coggins, NASA’s deputy associate administrator for SCaN at NASA Headquarters in Washington. “This data will help inform our transition to a commercial-first approach, ultimately strengthening the infrastructure needed to support Artemis missions and our long-term Moon to Mars objectives.” 

    Responses are due by 5 p.m. EDT on Monday, Oct. 27.

    NASA’s SCaN Program serves as the management office for the agency’s space communications and navigation systems. More than 100 NASA and non-NASA missions rely on SCaN’s two networks, the Near Space Network and the Deep Space Network, to support astronauts aboard the International Space Station and future Artemis missions, monitor Earth’s weather, support lunar exploration, and uncover the solar system and beyond.

    Artemis II will help confirm the systems and hardware needed for human deep space exploration. This mission is the first crewed flight under NASA’s Artemis campaign and is another step toward new U.S.-crewed missions on the Moon’s surface that will help the agency prepare to send American astronauts to Mars.

    Continue Reading

  • Iceland’s Mantle Plume Scatters Ancient North Atlantic Volcanoes

    Iceland’s Mantle Plume Scatters Ancient North Atlantic Volcanoes

    What do the rumblings of Iceland’s volcanoes have in common with the now peaceful volcanic islands off Scotland’s western coast and the spectacular basalt columns of the Giant’s Causeway in Northern Ireland?

    About sixty million years ago, the Icelandic mantle plume—a fountain of hot rock that rises from Earth’s core-mantle boundary—unleashed volcanic activity across a vast area of the North Atlantic, extending from Scotland and Ireland to Greenland.

    For decades, scientists have puzzled over why this burst of volcanism was so extensive. Now, research led by the University of Cambridge has found that differences in the thickness of tectonic plates around the North Atlantic might explain the widespread volcanism.

    The researchers compiled seismic and temperature maps of Earth’s interior, finding that patches of thinner tectonic plate acted like conduits, funnelling the plume’s molten rock over a wide area.

    Iceland , which is one of the most volcanically active places on Earth, owes its origin largely to the mantle plume. Beyond volcanism, the Iceland Plume’s influence even extends to shaping the seafloor and ocean circulation in the North Atlantic and, in turn, climate through time. Despite its global significance, many aspects of the plume’s behaviour and history remain elusive.

    “Scientists have a lot of unanswered questions about the Iceland plume,” said Raffaele Bonadio , a geophysicist at Cambridge’s Department of Earth Sciences and lead author of the study.

    Bonadio set out to explain why the plume’s volcanic imprint was much more widespread sixty million years ago—before the Atlantic opened—forming volcanoes and lava outpourings stretching over thousands of kilometres. The pattern could be explained by the mantle plume spreading outward in a branched, flowing formation, Bonadio explained, “but evidence for such flow has been scarce.”

    In search of answers, Bonadio focussed on a segment of the North Atlantic Igneous Province to better understand the complex distribution of volcanoes in Scotland and Ireland. He wanted to know if the structure of Earth’s tectonic plates played a role in the surface expression of volcanism.

    Using seismic data extracted from earthquakes, Bonadio created a computer-generated image of Earth’s interior beneath Britain and Ireland. This method, known as seismic tomography, works similarly to a medical CT scan, revealing hidden structures deep within the planet. Bonadio coupled this with seismic thermography measurements—a new method developed by the team—which reveal variations in the temperature and thickness of the tectonic plate.

    He found that northwest Scotland and Ireland’s volcanoes formed in areas where the lithosphere (Earth’s rigid outer layer that makes up the tectonic plates) is thinner and weaker.

    “We see ancient volcanoes concentrated within this corridor of thin lithosphere beneath the Irish Sea and surrounding areas,” said Bonadio. He thinks the hot plume material was preferentially funnelled along this corridor, ponding in the thin plate areas due to its buoyancy.

    Previously, some scientists had put forward alternative, non-mantle plume origins for the volcanic activity, said Bonadio. But his new research shows the scattering could be explained by the magma being diverted and re-routed to areas of thinner lithosphere.

    Sergei Lebedev, from the University of Cambridge said, “this striking correlation suggests that hot plume material eroded the lithosphere in this region. This resulting combination of thin lithosphere, hot asthenosphere and decompression melting likely caused the uplift and volcanic activity.”

    Previously, the authors have found a close link between the uneven distribution of earthquakes in Britain and Ireland and the thickness of the lithosphere, showing how the scars left by the mantle plume influence seismic hazards today.

    Bonadio and Lebedev are also using their methods to map geothermal energy resource potential. “In Britain and Ireland, the greatest supply of heat from the Earth’s mantle is in the same places where volcanoes erupted sixty million years ago, and where the lithosphere is thinner,” said Lebedev. He and Bonadio are working with international colleagues to apply their new seismic thermography methods to global geothermal assessment.

    /Public Release. This material from the originating organization/author(s) might be of the point-in-time nature, and edited for clarity, style and length. Mirage.News does not take institutional positions or sides, and all views, positions, and conclusions expressed herein are solely those of the author(s).View in full here.

    Continue Reading