More stories

  • in

    Respiratory bacteriome and its predicted functional profiles in blue whales (Balaenoptera musculus)

    AbstractThe respiratory microbiome plays a critical role in the health of organisms and studying it in natural populations can reveal interactions between hosts and their environment, as well as help predict responses to environmental stressors. We characterized the core respiratory bacteriome and functional profiles of Eastern North Pacific blue whales (Balaenoptera musculus) sampled in the Gulf of California using next-generation sequencing. Our compositional analysis identified 15 dominant bacterial phyla in the respiratory tract, with Proteobacteria (34.44%), Firmicutes (26.98%), Bacteroidota (20.26%), Fusobacteriota (7.61%), and Actinobacteria (5.55%) as the most abundant. Nineteen ASVs, representing 12 bacterial genera (primarily Corynebacterium, Oceanivirga, Tenacibaculum, and Psychrobacter), were shared by over 60% of whales, with a relative abundance greater than 0.02%. These bacteria, proposed to be the core respiratory bacteriome of blue whales, contributed to functional pathways associated with metabolism, environmental information processing, and cellular processes. Notably, two whales with high relative abundance of Mycoplasma spp. and of Streptococcus spp., exhibited overrepresented pathways related to nucleotide metabolism and translation, suggesting a suboptimal immune status or dysbiosis. To our knowledge, this is the first functional profiling of the bacteriome in any cetacean. Future studies are needed to explore how the blue whale respiratory bacteriome may vary over time, seasonally or across geographical locations. This study establishes a baseline for future research on the plasticity of the bacteriome, its associations with other microbiome components, the impact of environmental changes on its diversity, and its relevance for health. Our novel approach underscores the ecological and physiological importance of the bacteriome and its potential for long-term monitoring of a sentinel marine species in a rapidly changing ocean.

    Similar content being viewed by others

    Discovery of a novel bacterial class with the capacity to drive sulfur cycling and microbiome structure in a paleo-ocean analog

    Article
    Open access
    18 August 2023

    Role of the upper airway microbiota in respiratory virus and bacterial pathobiont dynamics in the first year of life

    Article
    Open access
    04 June 2025

    Phage resistance mutations in a marine bacterium impact biogeochemically relevant cellular processes

    Article

    04 December 2025

    IntroductionThe advent of modern technologies that allow for the identification of bacteria in environmental or clinical samples1 has led to a surge in studies examining the abundance, diversity, and structure of microbiomes across species2. Increasing our understanding of the microbiome is crucial because microbial communities associated with specific organs or tissues can significantly impact host physiology3 and health4. For instance, respiratory infections may arise when opportunistic microorganisms—normally part of a healthy respiratory tract—proliferate under certain conditions1, disrupting the diversity and composition of the microbial community in a phenomenon known as dysbiosis5, which can contribute to disease. Additionally, respiratory disease can result from exposure to non-commensal microorganisms with pathogenic potential. This underscores the importance of microbiome composition as a potential predictor of health and disease progression, often more so than the mere presence of specific microorganisms commonly associated with disease. Understanding how microbiomes differ between individuals could, therefore, become a valuable tool for assessing health6.When using the microbiome to assess health status, it is important to distinguish between commensal, opportunistic, and transient bacteria7. This distinction is complex, as the symbiotic relationships of bacteria can vary both across species and among individuals8 To help differentiate potential commensal and mutualistic bacteria, it is necessary to identify the core microbiome—the microbial taxa that predominate within a community and are common in apparently healthy individuals9. Defining the core bacteriome (the bacterial community of the microbiome) involves setting the detection threshold (relative abundance) and determining the minimum occurrence percentage (prevalence) of bacterial taxa to include10. However, because biological justifications for these prevalence and threshold values are often lacking11, it is important to exercise caution when interpreting results10. Despite varying definitions, the core bacteriome tends to be relatively stable, particularly when samples from closely related individuals are analyzed11,12.Microbial taxonomic composition provides a basic understanding of the microbiome, but it does not fully capture the intricate microbial contributions to host health13. Bacteria within the mammalian microbiome exist in composite communities1, whose diversity and abundance result from complex interactions between species14. The metabolic contributions of these communities are important to the host and depend on their composition15. This is where functional profiling becomes important, as it reveals the metabolic and ecological roles of microbial communities16,17. Combined taxonomic and functional analyses offer a deeper understanding of the microbiome’s dual nature, as both a diverse community and a functional unit that essential for the holobiont’s processes18. Functional predictions are based on genetic data derived from sequencing the 16 S rRNA gene18,19, which is mapped to reference databases, correlating specific taxa with known functional abilities20. These functions are organized hierarchically, from broad functional categories to specific metabolic pathways21. This approach allows researchers to infer the ecological and metabolic roles of the microbiome without the need for whole-genome sequencing, making it a powerful and accessible tool for microbiome research19. By integrating taxonomic and functional analyses, we can gain deeper insights into a host’s microbiome and its role in holobiont resilience, particularly in the context of health22.In cetaceans, characterizing the microbiome offers a unique opportunity to link microbial community structure and function with host ecology, physiology, and responses to environmental change, providing valuable insights for conservation and health monitoring23,24. Whales, as long-lived animals that play a critical role in the ocean’s carbon movement and storage, are vital to marine ecosystems25, and are often considered sentinels of ocean health26. The study of the cetacean microbiome is still in its early stages. Microbial diversity has been assessed for a few species1,27,28,29,30,31, and some opportunistic pathogens in the respiratory tracts of free-ranging cetaceans have also been described32,33. However, to our knowledge, no study has yet combined taxonomic and functional profiling of the microbiome in any cetacean species. Blue whales, among the world’s largest and most iconic animals, play an essential role in marine ecosystems34. Their long migrations and diverse habitats make them valuable indicators of ocean health35. Despite this, only one published study has examined the respiratory microbiome of blue whales in the wild36. Given the growing importance of understanding blue whale health in their natural environment, studying their respiratory microbiome is both timely and relevant. Not only would it provide insights into their exposure with potential pathogens32, but it would also establish a baseline of core bacteria and functional profiles in apparently healthy individuals. This baseline could facilitate the identification of dysbiosis, help predict potential diseases and ultimately inform conservation strategies and management plans for the species37. This is a pressing need, especially in light of the global and local environmental changes currently affecting oceans38. Here, we characterized the common core and functional profiles of the respiratory bacteriome in Eastern North Pacific blue whales from the Gulf of California using next-generation sequencing on blow samples collected from 17 adult blue whales via a non-invasive drone-based technique39.ResultsA total of 19 samples were analysed, including 17 photo-identified blue whales, one technical control, and one seawater sample. Exhaled breath was collected from the whales using a drone-based method previously described39, with no adverse behavior observed before, during, or after sampling. After filtering, denoising, merging, and chimera elimination (2.38% of reads), we obtained 68,922 sequences (mean per sample: 3514.8 [SD = 1998.3]), which corresponded to 1304 amplicon sequence variants (ASVs). We removed 51 ASVs classified as Archaea (n = 2), chloroplasts (n = 27), or mitochondria (n = 7), as well as those not classified at the phylum level (n = 15), and 22 ASVs identified as contaminants using the Decontam algorithm based on the LabControl sample reads. This left 1231 ASVs remained, with 500 ASVs classified as “Others” (representing less than 0.02% relative abundance).Species richness (S) in the blow samples ranged from 62 to 404 (mean = 189.63.06 [SD = 113.71]), and Simpson’s diversity index (D) ranged from 0.49 to 0.98 (mean = 0.94 [SD = 0.11]). The compositional analysis identified 15 bacterial genera (Fig. 1, Table S1) with Psychrobacter spp. (mean = 12.07% [SD = 6.09%]), Oceanovirga spp. (mean = 10.93% [SD = 4.42%]), Tenacibaculum spp. (8.87% [SD = 6.39%]), and Streptococcus spp. (6.79% [SD = 20.13%]) being the most abundant. Notably, two blow samples Bm057 and Bm044) exhibited a high relative abundance of the opportunistic pathogens Mycoplasma spp. (27.22%), and Streptococcus spp. (74.37%). In addition, we identified Bacteroides sp. in sample Bm042 at a high relative abundance of 13.96% compared with the other samples (mean = 0.02% [SD = 0.04%]). For this whale, mucus was also retrieved during blow sampling, which had a noticeable bad smell and a yellowish coloration; features that were not observed in any other samples.Fig. 1Stacked bar plot depicting relative abundance of the top 15 bacterial genera. Each vertical bar depicts the relative abundance of adjusted sequence variants (ASVs) and associated taxa that were recovered per sample. Plot shows the top fifteen identified bacterial genera, unclassified, and “others” (sum of bacteria that did not reach the detection threshold of 0.02%).Full size imageFunctional profiling, at 97% similarity, was possible for 32.53% of the ASVs. At KEGG Level 1, the most predominant pathways were associated with metabolism (mean = 74.88% [SD = 4.36%]), followed by environmental information processing (mean = 9.98% [SD = 1.82%]), cellular processes (mean = 5.54% [SD = 1.67%]), and genetic information processing (mean = 5.40% [SD = 1.57%]). At KEGG Level 2, the top subcategories included global and overview maps (mean = 38.37% [SD = 2.96%]), carbohydrate metabolism (mean = 9.65% [SD = 1.14%]), amino acid metabolism (mean = 7.23% [SD = 1.45%]), and membrane transport (mean = 6.55% [SD = 1.32%]). At KEGG Level 3, the most abundant pathways were metabolic pathways, biosynthesis of secondary metabolites, ABC transporters, and microbial metabolism in diverse environments (Fig. 2).Fig. 2Alluvial diagram of the top 20 predicted functional pathways (at different KEEG levels) associated with the bacteriome in the respiratory tract of blue whales.Full size imageIn two blow samples (Bm057 and Bm044) with the highest relative abundance of opportunistic pathogens, functional profiling revealed overrepresentation of pathways such as nucleotide metabolism, membrane transport, translation, folding, sorting and degradation, and carbohydrate metabolism; while pathways related to amino acid metabolism, cofactor and vitamin metabolism, lipid metabolism, and biosynthesis of secondary metabolites were underrepresented (Fig. 3). Among the bacterial genera identified, Psychrobacter (26.83%) contributed most to the functional pathways predicted in the blue whale respiratory tract, followed by Tenacibaculum (17.48%) and Porphyromonas (13.01%). Genera such as Suttonella and Streptococcus contributed less (4.07% and 3.25%, respectively; Fig. S1). Despite variation in taxonomic composition, functional profiles across individuals were consistent (Fig. S2).Fig. 3Boxplot of the relative abundance of functional pathways (at KEGG Level 2) across all blow samples. Red dots represent blow sample Bm057 (the whale that had a high relative abundance of Mycoplasma sp.), while yellow dots correspond to blow sample Bm042 (the whale that had a high relative abundance of Streptococcus sp.). Functional pathways that were overrepresented in both Bm057 and Bm044 compared to all other samples (grey) are highlighted as light green columns, while underrepresented pathways are shown as light blue columns. The functional pathway that was underrepresented in the bacteriome of Bm057 but over represented in Bm042 is highlighted as light purple columns, and the functional pathways were over represented only in Bm057 are shown in as light orange columns.Full size imageThe core bacteriome analysis identified 19 ASVs from 12 bacterial families (Fig. 4, Table S2), with Tenacibaculum (ASV3) and Oceavivirga (ASV7) the being the most abundant genera (30.01% [SD = 15.94] and 28.78% [SD = 9.76], respectively]). The core functional profile derived from these core ASVs was composed mainly by metabolic pathways (24.99%), biosynthesis of secondary metabolites (11.09%), biosynthesis of antibiotics (8.73%), and microbial metabolism in diverse environments (8.34%) (KEEG level 1; Fig. S3).Fig. 4Relative abundances of bacterial genera that constitute the core respiratory bacteriome of the blue whale. The figure includes the seven ASVs that were present in more than 60% of the samples and that had a relative abundance of over 0.02%. The relative abundance of each ASV shown in this plot is confined to the core microbiome members and not the entire microbiome of each sample.Full size imageThe Bayesian approach used to estimate the contribution of seawater diversity to blow samples indicated that seawater contributed on average of 1.68% (SD = 0.81). Herbaspirillum sp., the most common genus in seawater (20.23% relative abundance; Fig. 1), was also detected in blow samples, albeit at a lower average abundance (3.39%; SD = 5.97). Interestingly, three whale blows (from individuals Bm023, Bm043, and Bm059) exhibited notably higher levels of Herbaspirillum sp. (9.44%, 15.65%, and 16.82%, respectively).DiscussionA healthy microbiome is generally characterized by high diversity, which helps both the microbiome and the host cope with external challenges30. In our study, the respiratory bacteriome of the blue whale exhibited considerable diversity, with significant variation in bacterial richness and abundance across samples. These fluctuations may arise from several factors, including bacterial immigration from the environment during inhalation, mucociliary clearance, and community growth rates40, all of which can vary among healthy individuals27. However, variations could also stem from sampling techniques, such as differences in the number of blows, volume of sample collected, whale size and behavior (e.g. dive depth and duration) 39,41. Notably, the bacterial diversity observed in blue whale blow samples was similar to that reported for humpback whales and bottlenose dolphins41,42, although the blue whale blow showed greater taxonomic richness. This may be attributed to differences in methods used to resolve taxonomy43,44,45 or the identification of rare bacterial species44,45, which play an important role in microbiome resilience, given their contribution as a seed bank of genetic resources that can lead to the restoration of the core microbiome46.The presence of a complex respiratory bacteriome is beneficial for a host, as higher microbial diversity supports vital ecosystem functions47. Functional analysis of the blue whale bacteriome revealed overrepresentation of pathways related to macromolecular metabolism and environmental information processing and signal transduction, indicating a potential role in adapting to environmental changes48,49. This result reinforces the idea that bacteriome diversity serves a protective role for the host18, as these pathways are critical for maintaining host health and epithelial immune function50,51 by enabling microbial communication with host immune cells via molecular signals that activate pattern recognition receptors, triggering cytokine production and immune cell recruitment52, including dendritic cells53.Our findings indicate that the respiratory bacteriome of blue whales is dominated by members of Proteobacteria, Firmicutes, Bacteroidota, Actinobacteria, and Fusobacteriota, which are common bacterial phyla in the respiratory microbiome of other mammals1. Particularly noteworthy is the consistent presence of Psychrobacter sp. and Tenacibaculum sp., which are known commensal bacteria54,55 that contribute to respiratory and skin health4,27,29,30,56, although they can also be implicated in pathological conditions in other organs57,58. Additionally, the respiratory core bacteriome included Oceanivirga sp., a bacterium common to the respiratory tract of various marine mammals from different geographical locations59, and identified as part of the core respiratory bacteriome of humpback whales41. Given that Oceanivirga sp., was present in most of the blue whales sampled, it is reasonable to consider it a key member of their respiratory bacteriome, reflecting a healthy respiratory epithelium.It is important to recognize that while the bacterial taxa in the blue whale’s respiratory bacteriome share similarities with those found in the oropharynx and nasopharynx of terrestrial mammals60, cetaceans lack anatomical connections between the mouth and nasopharynx41. Thus, the bacteria identified in this study are more likely associated with the respiratory tract rather than the oral cavity of the blue whales. In addition, it is important to note that this composition may vary over time and space, and could be influenced by factors such as fasting, reproductive stage6,61, or other physiological variables1,62,63.Interestingly, four bacterial genera (Psychrobacter, Tenacibaculum, Staphylococcus, and Corynebacterium) identified in the blow samples are typically found in the skin of humans and other terrestrial mammals64,65. These genera were also identified in the skin microbiota of both captive and free-ranging cetaceans1,6,28,55,56. Given that strict protocols were followed to minimize contamination during sampling, processing or sequencing, their presence in whale blow suggests that they colonize the epithelial lining of the blowhole and are forcefully expelled during exhalation41. Moreover, Psychrobacter and Tenacibaculum, contributed significantly to metabolic and environmental processing pathways, suggesting their role in maintaining microbial and host homeostasis. We hypothesize that these bacteria establish a commensal or mutualistic associations with the blue whale, potentially offering a protective role against dysbiosis and environmental stressors. Furthermore, it is possible that these taxa play a crucial role in maintaining respiratory health in this species, and more detailed functional analyses will be necessary in the future to clarify their ecological and physiological roles.Our study also found that approximately 2% of the microbial diversity in blow samples overlapped that of seawater, indicating some influence of the marine environment on the respiratory bacteriome, possibly as carryover during diving immersions. However, this overlap should be interpreted with caution, as seawater sampling was limited in number and not conducted for every breath sample. The absence of more water samples restricts our ability to fully assess the extent to which environmental microorganisms contribute to the respiratory bacteriome composition. Regardless, the detection of Psychrobacter, Oceanivirga, Tenacibaculum, Helcococcus, Porphyromonas, Mycoplasma, Dielma, Synechococcus, and Suttonella, in blue whale blow, but not in seawater, adds support to the notion that these taxa are intrinsic to the blue whale’s respiratory microbiome. Variations in the relative abundance of Herbaspirillum sp. in certain samples suggest that whale diving behavior, environmental factors and technical sampling conditions may also influence bacterial detection.We identified Bacteroides spp. in blow Bm042 at a relative abundance of 13.96%. Bacteroides spp. can influence airway immune responses by inducing regulatory T cells and associated cytokines and has been shown to promote transient PD-L1 expression and modulate general aeroallergen responses66. This genus has also been reported in increased abundance during tracheobronchitis, suggesting potential roles in modulating respiratory immune function66,67. Interestingly, whale Bm042 also presented mucus with a yellowish coloration, which may indicate a high concentration of airway mucin, which is associated with various pulmonary diseases68. The excessive synthesis of mucin can result from increased neutrophil recruitment, reflecting an acute inflammatory response to bacterial infection in the airways69,70. Given its immunomodulatory capacity, the elevated abundance of Bacteroides spp. in the blow of whale Bm042 may reflect a role in host immune regulation during localized airway infection or inflammation. Its presence alongside signs of mucus suggests a potential microbial shift and underscores the need to consider both protective and pathogenic roles of Bacteroides spp. in the respiratory tract.Two unidentified species from Mycoplasma and Streptococcus were found in the blow of two whales. As 16 S rRNA gene sequencing does not allow reliable species-level resolution, our assignments were limited to the genus level, and we acknowledge that the detected Streptococcus and Mycoplasma taxa may include both commensal and opportunistic members. This taxonomic uncertainty underscores the importance of continued monitoring, since shifts at the genus level can still provide meaningful indicators of host health. Various species within these bacterial genera are known respiratory tract opportunists in mammals71,72 and have been detected in the lungs of stranded marine mammals73,74, although their presence does not necessarily indicate disease since they can also occur in healthy hosts72. This is essential to consider when studying the bacteriome of individuals, as the type of relationship between host and bacteria can depend on different factors, including the status of the immune system8,71. The low prevalence of these pathogens in our study likely suggests that they are not common members of the respiratory bacterial community and highlights the natural diversity of the blue whale respiratory microbiome. As the blue whales migrate through coastal areas, they could become exposed to transient bacteria which do not normally manage to colonize the respiratory epithelium. However, the intense maritime traffic and potential human interactions75 could act as stressors that affect immune regulation of bacterial communities in susceptible hosts and favor the growth of transient or opportunistic bacteria68,69,70,76,77. Therefore, it is plausible that the detection of these bacteria could indicate underlying health conditions, a suboptimal immune status, or chronic stress in these individuals78. We have some support for this argument as the respiratory bacteriome of the two whales that harbored Mycoplasma spp. and Streptococcus spp. exhibited distinct functional pathway patterns than the other whales, whose bacteriome functional profiles remained largely stable across individuals. Namely the bacteriome of these two whales showed overexpression of nucleotide metabolism, translation, and replication and repair pathways, which have been associated with various diseases in humans79. In contrast, pathways involved in lipid metabolism and biosynthesis of other secondary metabolites were underrepresented in these whales, suggesting possible vulnerabilities in their immune responses, as has been shown for humans50,80. As these functional profiles were inferred from 16 S rRNA gene data, incorporating functional analyses based on transcriptomics or other omics approaches in future studies would provide a more comprehensive understanding of the microbiome’s functional potential. The identification of these bacterial genera and the distinct functional profile of the bacteriome of the whales that harbored them, highlights the need for ongoing monitoring specific microbial taxa, regardless of their perceived roles as commensal, mutualistic or opportunistic in other mammals, and underscores the importance of considering natural fluctuations in the respiratory bacteriome when assessing the health of blue whales.Given the current threats facing marine ecosystems, that include habitat degradation, pollution, and other anthropogenic stressors26, the taxonomic and functional study of the blue whale respiratory bacteriome offers valuable insights into their health and resilience. Respiratory microbiome data can serve as an early warning system by detecting shifts associated with environmental change, disease, or human activities41. Monitoring such changes in bacterial composition and functionality over time can help inform conservation efforts and management strategies23 to protect these iconic species and the ecosystems they inhabit. While our study is based on a modest number of individuals, it represents a meaningful fraction of the population migrating through the Loreto area. Future studies incorporating multiple blow samples per individual could capture temporal variability more effectively, reduce potential sampling bias, and further strengthen the value of microbiome monitoring for conservation and health assessments.Methods Sample collection Using a small Phantom 3® quadrocopter drone (DJI Innovations, China) with floaters and sterile Petri dishes, we collected 17 blows samples from 17 different individual blue whales sampled between February and March 2016 and 2017 in Loreto Bay National Park (25° 51′ 51″ N, 111° 07′ 18″ O) within the Gulf of California, Mexico. The number of sampled whales represents 17% of the estimated 100 blue whales that reside during winter/spring in the southwestern Gulf of California (mark-recapture data from 1994 to 200681. Each whale was photo-identified prior to sample collection81. The approach of the drone to the whale was done from the caudal fin towards the head to minimize disturbance, and sampling was conducted at a height between 3 and 4 m above the blowhole39. We observed the whale body condition (see Supplementary Material) for each individual and recorded characteristics of their blow, such as color and odor when we were sufficiently close to the whale during sampling.For each sample, blow droplets were swabbed directly from the Petri dish using one sterile cotton-tipped swab per individual. These were then transferred to a sterile 1.5 mL cryogenic microtube containing 500 µL of 96% molecular grade ethanol and kept frozen in a liquid nitrogen container until processing. To address potential contamination, all necessary precautions were taken, always including the use of sterile gloves and face masks during sample processing. In addition, we included a technical control, termed “LabControl” (a template-free DNA negative extraction control), to identify any contaminants during sample processing. Furthermore, we included a seawater sample, termed “seawater” (a DNA sample extracted from 1mL water collected at a depth of 0.10 m in the same area where we sampled the whale blows), to consider potential sources of bacterial diversity for the blow samples. DNA extraction, PCR amplification and sequencing Total DNA was isolated from the whale blow, seawater, and LabControl samples in one batch using a QIAamp ® DNA Mini Kit (QIAGEN, Germany). The primers used for sequencing the 16S rRNA V3 and V4 regions were 341F (5′-CCTACGGGNGGCWGCAG) and 785R (5′-GACTACHVGGGTATCTAATCC), which amplified a single product of 444 bp82. The PCR program used an initial denaturation step at 95 °C for 3 min; 25 cycles of 95 °C for 30 s, 55 °C for 30 s, and 72 °C for 30 s; and a final extension step at 72 °C for 5 min. Each 25 µL-reaction contained 12.5 ng of extracted DNA, 5 µM of barcoded primers and 2x KAPA HiFi HotStart Ready Mix (KAPABIOSYSTEM, Cape Town, South Africa). 1 µl of each sample was run on a 2100 Bioanalyzer (Agilent Technologies, CA, USA) with an Agilent DNA 1000 chip (Agilent Technologies, CA, USA) to verify amplicon size. AMPure XP beads (New England BioLabs, USA) were used to remove unused primers and primer dimers. Amplicons were sequenced over 2- by 250-bp MiSeq at the Unit of Sequencing and Identification of Polymorphisms of the National Institute of Genomic Medicine (Instituto Nacional de Medicina Genómica, Unidad de Secuenciación e Identificación de Polimorfismos, INMEGEN) in Mexico. Dual index barcodes were used to avoid index hopping83. The protocol used by INMEGEN can be seen in: https://support.illumina.com/documents/documentation/chemistry_documentation/16s/16s-metagenomic-library-prep-guide-15044223-b.pdf. 16 S rRNA sequence data processing A quality control overview was performed using FASTQC84. This allowed us to obtain a quick impression of the data and avoid downstream problems. The raw sequences were then imported into R v.4.2.185, where all subsequent analyses were carried out. We used the Divisive Amplicon Denoising Algorithm 2 (dada2) v.1.26.044 to infer exact ASVs. This approach is preferable over the rough and less precise 16 S rRNA OTU clustering approach86 that groups the sequences with a 97% identity87. First, we filtered by quality (trunQ = 25) and discarded the sequences that presented more than two Ns (maxN = 0) or more than two expected errors (maxEE = 2). Next, the forward and reverse reads for each sample were combined into a single merged contig sequence, and we grouped all identical reads into unique sequences to determine their abundance. After building the ASVs table and removing chimeras (detected using self-referencing), sequences were classified and identified with Decipher v.2.26.088, using the SILVA rRNA sequence database v.138.1 as the taxa reference89. We used phyloseq v.1.42.090 to classify and remove any sequence not classified at the kingdom and Phylum level or belonging to Archaea, Eukarya, chloroplasts, or mitochondria. Contamination assessmentAt present, there is no standard approach for minimizing or controlling potential contaminants in 16 S rRNA gene sequencing experiments91. In our study, we employed two methods to limit and eliminate contaminant sequences from downstream analyses. First, we used metagMisc v.0.5.092 to eliminate ASVs with less than ten reads (minabund = 1093). Next, we used Decontam version 1.18.094 to identify sequences that had a negative relationship with DNA concentration. We classified ASVs found in the LabControl sample as potential contaminants if they were identified as true contaminants by the Decontam algorithm. To ensure result accuracy, we then removed the identified contaminant sequences from the analysis. Respiratory bacteriome analysis and identification of functional pathways To get a sense of the bacterial community composition of the samples, we used phyloseq to identify the distribution of read counts from all the samples and to plot the relative abundance stacked bar plot at genus level. In addition, we used SourceTracker95, a Bayesian approach that allowed us to estimate the proportion of the bacterial community in the blue whale blows samples that are also detected in the seawater sample. Using microbiome v.1.2096, we identified the common core bacteriome (threshold detection set at ≥ 0.02%, prevalence set at ≥ 60%). We selected these values because we wanted a more conservative approach. Finally, we calculated alpha diversity indices: richness (S) and Simpson’s diversity index (D) using vegan v.2.6.497. Bacterial functional profiles and pathways were inferred from 16 S rRNA gene sequencing data and annotated at a 97% similarity threshold using the ref99NR database as a reference, employing the Tax4Fun2 package19, which is based on the Kyoto Encyclopedia of Genes and Genomes (KEGG; 20). All graphs were rendered using Tableau v.2024.398 and RAWGraphs v 2.099.Use of animals in research All methods were performed in accordance with the relevant international guidelines and regulations of Mexican authorities. See ethical approval.We confirm that our manuscript complies with the ARRIVE Essential 10 guidelines. The study design, experimental groups, and units are clearly described, with exact sample sizes reported. All outcome measures, including the primary outcome, are clearly defined. Statistical methods and assumptions are detailed, along with the software used. Comprehensive information on the animals, including species and probable health status, is provided. Experimental procedures are described with sufficient detail to allow replication, including what was done, how, when, where, and why. Results are presented with descriptive statistics and measures of variability, along with confidence intervals where appropriate.

    Data availability

    Data from the Sequence Read Archive (SRA) submission will be released upon publication. Accession ID: PRJNA977688.
    ReferencesRhodes, L. D., Emmons, C. K., Wisswaesser, G. S., Wells, A. H. & Hanson, M. B. Bacterial microbiomes from mucus and breath of Southern resident killer whales (Orcinus orca). Conserv. Physiol. 10, (2022).Watkins, C. A. et al. A comparative study of the fecal microbiota of Gray seal pups and yearlings—a marine mammal Sentinel species. Microbiol. Open. 11, e1281 (2022).Article 
    CAS 

    Google Scholar 
    Foster, K. R., Schluter, J., Coyte, K. Z. & Rakoff-Nahoum The evolution of the host microbiome as an ecosystem on a leash. Nature 548, 43–51 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Bierlich, K. C. et al. Temporal and regional variability in the skin Microbiome of humpback whales along the Western Antarctic Peninsula. Appl. Environ. Microbiol. 84, e02574–e02517 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sehnal, L. et al. Microbiome composition and function in aquatic vertebrates: small organisms making big impacts on aquatic animal health. Front. Microbiol. 12, 567408 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dominguez-Sanchez, C. A., Ferguson, S. H., Edkins, T., Young, B. G. & Kringorn, J. Pilot study: decoding the skin Microbiome of bowhead (Balaena mysticetus) and killer whales (Orcinus orca) in Nunavut, Canada. Arct. Sci. 10, 169–188 (2024).Article 

    Google Scholar 
    Infante-Villamil, S., Huerlimann, R. & Jerry, D. R. Microbiome diversity and dysbiosis in aquaculture. Rev. Aquac. 13, 1077–1096 (2021).Article 

    Google Scholar 
    Wiesmann, C. L., Wang, N. R., Zhang, Y., Liu, Z. & Haney, C. H. Origins of symbiosis: shared mechanisms underlying microbial pathogenesis, commensalism and mutualism of plants and animals. FEMS Microbiol. Rev. 47, fuac048 (2022).Article 
    PubMed Central 

    Google Scholar 
    Willis, A., Bunge, J. & Whitman, T. Improved detection of changes in species richness in high diversity microbial communities. J. R. Stat. Soc. Ser. C Appl. Stat. 66, 963–977 (2017).Article 
    MathSciNet 

    Google Scholar 
    Neu, A. T., Allen, E. E. & Roy, K. Defining and quantifying the core microbiome: Challenges and prospects. Proc. Natl. Acad. Sci. USA 118, (2021).Risely, A. Applying the core Microbiome to understand host–microbe systems. J. Anim. Ecol. 1–10 (2020).Vendl, C., Nelson, T., Ferrari, B., Thomas, T. & Rogers, T. Highly abundant core taxa in the blow within and across captive bottlenose dolphins provide evidence for a temporally stable airway microbiota. BMC Microbiol. 1–29 (2020).Louca, S. et al. Function and functional redundancy in microbial systems. Nat. Ecol. Evol. 2, 936–943 (2018).Article 
    PubMed 

    Google Scholar 
    Stubbendieck, R. M., Vargas-Bautista, C. & Straight, P. D. Bacterial communities: interactions to scale. Front. Microbiol. 7, 1–19 (2016).Article 

    Google Scholar 
    Turnbaugh, P. J. et al. A core gut microbiome in obese and lean twins. Nature 457, 480–484 (2008).Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Toole, D. R., Zhao, J., Martens-Habbena, W. & Strauss, S. L. Bacterial functional prediction tools detect but underestimate metabolic diversity compared to shotgun metagenomics in Southwest Florida soils. Appl. Soil. Ecol. 168, 104123 (2021).Article 

    Google Scholar 
    Grisnik, M., Grinath, J. B., Munafo, J. P. & Walker, D. M. Functional redundancy in Bat microbial assemblage in the presence of the white nose pathogen. Microb. Ecol. 86, 713–726 (2023).Article 
    CAS 
    PubMed 
    ADS 

    Google Scholar 
    Zhang, Y., Chen, R., Zhang, D. D., Qi, S. & Liu, Y. Metabolite interactions between host and microbiota during health and disease: which feeds the other? Biomed. Pharmacother. 160, 114295 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wemheuer, F. et al. Tax4Fun2: prediction of habitat-specific functional profiles and functional redundancy based on 16S rRNA gene sequences. Environ. Microbiomes. 15, 1–12 (2020).
    Google Scholar 
    Kanehisa, M., Sato, Y. & Morishima, K. BlastKOALA and ghostkoala: KEGG tools for functional characterization of genome and metagenome sequences. J. Mol. Biol. 428, 726–731 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kanehisa, M., Furumichi, M., Sato, Y., Kawashima, M. & Ishiguro-Watanabe, M. KEGG for taxonomy-based analysis of pathways and genomes. Nucleic Acids Res. 51, D587–D592 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Li, J. et al. Bacterial community composition and function of tropical river ecosystem along the Nandu river on Hainan Island, China. Int. J. Environ. Res. Public. Health. 20, 382 (2023).Article 
    CAS 

    Google Scholar 
    West, A. G. et al. The microbiome in threatened species conservation. Biol. Conserv. 229, 85–98 (2019).Article 

    Google Scholar 
    Bahrndorff, S., Alemu, T., Alemneh, T. & Nielsen, J. L. The microbiome of animals: Implications for conservation biology. Int. J. Genomics 2016, 1–7 (2016).Pershing, A. J., Christensen, L. B., Record, N. R., Sherwood, G. D. & Stetson, P. B. The impact of whaling on the ocean carbon cycle: why bigger was better. PLoS One. 5, e12444 (2010).Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Palmer, E. et al. A piece of the puzzle: analyses of recent strandings and historical records reveal new genetic and ecological insights on new Zealand sperm whales. Mar. Ecol. Prog. Ser. 690, 201–217 (2022).Article 
    CAS 
    ADS 

    Google Scholar 
    Toro, F. et al. Composition and structure of the skin microbiota of Rorquals off the Eastern South Pacific. FEMS Microbiol. Ecol. 97, 1–10 (2021).Article 

    Google Scholar 
    Hooper, R. et al. Host-derived population genomics data provides insights into bacterial and diatom composition of the killer Whale skin. Mol. Ecol. 28, 484–502 (2019).Article 
    MathSciNet 
    PubMed 

    Google Scholar 
    Apprill, A. et al. Marine mammal skin microbiotas are influenced by host phylogeny. R Soc. Open. Sci. 7, 192046 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Van Cise, A. M. et al. Skin Microbiome of Beluga whales: spatial, temporal, and health-related dynamics. Anim. Microbiome. 2, 39 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vendl, C., Nelson, T., Ferrari, B., Thomas, T. & Rogers, T. Highly abundant core taxa in the blow within and across captive bottlenose dolphins provide evidence for a temporally stable airway microbiota. BMC Microbiol 21, (2021).Raverty, S. et al. Respiratory microbiome of endangered Southern resident killer whales and microbiota of surrounding sea surface microlayer in the Eastern North Pacific. Sci. Rep. 7, 1–12 (2017).Article 
    CAS 

    Google Scholar 
    Acevedo-Whitehouse, K. A., Rocha-Gosselin, A. & Gendron, D. A novel non-invasive tool for disease surveillance of free-ranging whales and its relevance to conservation programs. Anim. Conserv. 13, 217–225 (2010).Article 

    Google Scholar 
    Attard, C., Beheregaray, L. & Möller, L. Towards population-level conservation in the critically endangered Antarctic blue whale: the number and distribution of their populations. Sci. Rep. 6, 22291 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Rice, A. et al. Update on frequency decline of Northeast Pacific blue Whale (Balaenoptera musculus) calls. PLoS One 17, (2022).Vendl, C. et al. Does sociality drive diversity and composition of airway microbiota in cetaceans? Environ. Microbiol. Rep. 12, 324–333 (2020).Article 
    PubMed 

    Google Scholar 
    Raverty, S. et al. Pathology findings and correlation with body condition index in stranded killer whales (Orcinus orca) in the Northeastern Pacific and Hawaii from 2004 to 2013. PLoS One 15, (2020).Barlow, D. R., Klinck, H., Ponirakis, D., Branch, T. A. & Torres, L. G. Environmental conditions and marine heatwaves influence blue Whale foraging and reproductive effort. Ecol. Evol 13, (2023).Domínguez-Sánchez, C. A., Acevedo-Whitehouse, K. A. & Gendron, D. Effect of drone-based blow sampling on blue Whale (Balaenoptera musculus) behavior. Mar. Mamm. Sci. 34, 841–850 (2018).Article 

    Google Scholar 
    Huffnagle, G. B., Dickson, R. P. & Lukacs, N. W. The respiratory tract microbiome and lung inflammation: A two-way street. Mucosal Immunol. 10, 299–306 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Apprill, A. et al. Extensive core Microbiome in drone-captured Whale blow supports a framework for health monitoring. mSystems 2, e00119–e00117 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bik, E. M. et al. Marine mammals harbor unique microbiotas shaped by and yet distinct from the sea. Nat. Commun. 7, 1–13 (2016).Article 

    Google Scholar 
    Eren, A. M. et al. Minimum entropy decomposition: unsupervised oligotyping for sensitive partitioning of high-throughput marker gene sequences. ISME J. 9, 968–979 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Callahan, B. J. et al. DADA2: High-resolution sample inference from illumina amplicon data. Nat. Methods. 13, 581–583 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ahlgren, N. A., Perelman, J. N., Yeh, Y. C. & Fuhrman, J. A. Multi-year dynamics of fine-scale marine cyanobacterial populations are more strongly explained by phage interactions than abiotic, bottom-up factors. Environ. Microbiol. 21, 2948–2963 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Jousset, A. et al. Where less May be more: how the rare biosphere pulls ecosystems strings. ISME J. 11, 853–862 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Saraiva, J. P. et al. Mining synergistic microbial interactions: A roadmap on how to integrate multi-omics data. Microorganisms 9, 840 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Shaw, C., Hess, M. & Weimer, B. C. Two-component systems regulate bacterial virulence in response to the host Gastrointestinal environment and metabolic cues. Virulence 13, 1666–1680 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Liu, Y. et al. Transcriptomics reveals substance biosynthesis and transport on membranes of Listeria monocytogenes affected by antimicrobial lipopeptide brevilaterin B. Food Sci. Hum. Wellness. 12, 1359–1368 (2023).Article 

    Google Scholar 
    Silva, S. G. et al. Insights into the antimicrobial activities and metabolomes of Aquimarina (Flavobacteriaceae, Bacteroidetes) species from the rare marine biosphere. Mar. Drugs. 20, 423 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Du, J. et al. A decision analysis model for KEGG pathway analysis. BMC Bioinform. 17, 407 (2016).Article 

    Google Scholar 
    Wilden, J. J., Jacob, J. C., Ehrhardt, C., Ludwig, S. & Boergeling, Y. Altered signal transduction in the immune response to influenza virus and S. pneumoniae or S. aureus co-Infections. Int. J. Mol. Sci. 22, 5486 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wilson, K. R., Gressier, E., McConville, M. J. & Bedoui, S. Microbial metabolites in the maturation and activation of dendritic cells and their relevance for respiratory immunity. Front. Immunol. 13, 897462 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Centeno-Martinez, R. E., Klopp, R. N., Koziol, J., Boerman, J. P. & Johnson, T. A. Dynamics of the nasopharyngeal Microbiome of apparently healthy calves and those with clinical symptoms of bovine respiratory disease from disease diagnosis to recovery. Front. Vet. Sci. 10, 1297158 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Smith, S. A. et al. The respiratory microbiota of three cohabiting Beluga whales (Delphinapterus leucas) under human care. Front. Mar. Sci. 10, 1168623 (2023).Article 

    Google Scholar 
    Apprill, A. et al. Humpback Whale populations share a core skin bacterial community: towards a health index for marine mammals? PLoS One. 9, e90785 (2014).Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Li, C. et al. Insights on gut and skin wound microbiome in stranded Indo-Pacific finless Porpoise (Neophocaena phocaenoides). Microorganisms 10, (2022).Nowlan, J. P., Lumsden, J. S. & Russell, S. Advancements in characterizing tenacibaculum infections in Canada. Pathogens 9, 1029 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Volokhov, D. V. et al. Oceanivirga Miroungae sp. nov., isolated from oral cavity of Northern elephant seal (Mirounga angustirostris). Int. J. Syst. Evol. Microbiol. 70, 3037–3048 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Guglielmetti, S. et al. Oral bacteria as potential probiotics for the pharyngeal mucosa. Appl. Environ. Microbiol. 76, 3948–3958 (2010).Article 
    CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Gallego, J. C. Cetacean microbiota: ecological alterations and their impact on health of these mammals. SVOA Microbiol. 6, 66–70 (2025).Article 
    CAS 

    Google Scholar 
    Taneja, V. & Microbiome Impact of gender on function & characteristics of gut Microbiome. In Principles of Gender-Specific Medicine, 3rd ed., 569–583 (Academic, 2017).Björk, J. R., O’Hara, R. B., Ribes, M., Coma, R. & Montoya, J. M. The dynamic core microbiome: Structure, dynamics and stability. BioRxiv 137885. https://doi.org/10.1101/137885 (2018).Byrd, A. L., Belkaid, Y. & Segre, J. A. The human skin Microbiome. Nat. Rev. Microbiol. 16, 143–155 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Worthing, K. A. et al. Clonal diversity and geographic distribution of methicillin-resistant Staphylococcus Pseudintermedius from Australian animals: discovery of novel sequence types. Vet. Microbiol. 213, 58–65 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Gollwitzer, E. S. et al. Lung microbiota promotes tolerance to allergens in neonates via PD-L1. Nat. Med. 20, 642–647 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Li, R., Li, J. & Zhou, X. Lung microbiome: new insights into the pathogenesis of respiratory diseases. Signal. Transduct. Target. Ther. 9, 1–27 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Spies, R. et al. Sputum color as a marker for bacteria in acute exacerbations of chronic obstructive pulmonary disease: a systematic review and meta-analysis. Ann. Am. Thorac. Soc. 20, 738–748 (2023).Article 
    PubMed 

    Google Scholar 
    Stockley, R. A., O’Brien, C., Pye, A. & Hill, S. L. Relationship of sputum color to nature and outpatient management of acute exacerbations of COPD. Chest. 117, 1638–1645 (2000).Article 
    CAS 
    PubMed 

    Google Scholar 
    Meldrum, O. W. & Chotirmall, S. H. Mucus, microbiomes and pulmonary disease. Biomedicines 9, 675 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pereyre, S. & Tardy, F. Integrating the human and animal sides of Mycoplasmas resistance to antimicrobials. Antibiotics (Basel). 10, 1261 (2021).
    Google Scholar 
    Numberger, D., Siebert, U., Fulde, M. & Valentin-Weigand, P. Streptococcal infections in marine mammals. Microorganisms 9, 350 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Foster, G. et al. Mycoplasma species isolated from harbor porpoises (Phocoena phocoena) and a sowerby’s beaked Whale (Mesoplodon bidens) stranded in Scottish waters. J. Wildl. Dis. 47, 206–211 (2011).Article 
    PubMed 

    Google Scholar 
    Souter, R. et al. Fatal Streptococcus iniae infection in a juvenile free-ranging short-beaked common Dolphin (Delphinus delphis). Animals. 11, 3123 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Blevins, C. et al. Sex- and age-specific migratory strategies of blue whales in the Northeast Pacific ocean. Front. Mar. Sci. 9, 944918 (2022).Article 

    Google Scholar 
    Malmuthuge, N. et al. Effect of maternal separation and transportation stress on the bovine upper respiratory tract Microbiome and the immune response to resident opportunistic pathogens. Anim. Microbiome. 3, 1–18 (2021).Article 

    Google Scholar 
    Margolis, E., Yates, A. & Levin, B. R. The ecology of nasal colonization of Streptococcus pneumoniae, haemophilus influenzae and Staphylococcus aureus: the role of competition and interactions with host’s immune response. BMC Microbiol. 10, 1–11 (2010).Article 

    Google Scholar 
    Nicola, I. et al. Characterization of the upper and lower respiratory tract microbiota in Piedmontese calves. Microbiome 5, 152 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mullen, N. J. & Singh, P. K. Nucleotide metabolism: a pan-cancer metabolic dependency. Nat. Rev. Cancer. 23, 275 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Brown, E. M., Clardy, J. & Xavier, R. J. Gut Microbiome lipid metabolism and its impact on host physiology. Cell. Host Microbe. 31, 173–186 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gendron, D. & de la Ugalde, A. A new classification method to simplify blue Whale photo- identification technique. J. Cetacean Res. Manag. 12, 79–84 (2012).Article 

    Google Scholar 
    Thijs, S. et al. Comparative evaluation of four bacteria-specific primer pairs for 16S rRNA gene surveys. Front. Microbiol. 8, 494 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hornung, B. V. H., Zwittink, R. D. & Kuijper, E. J. Issues and current standards of controls in Microbiome research. FEMS Microbiol. Ecol. 95, fiz045 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Andrews, S. FastQC: A quality control tool for high throughput sequence data [Online]. (2010). http://www.bioinformatics.babraham.ac.uk/projects/fastqcR Core Team. R: A Language and Environment for Statistical Computing. [Online]. (2024). https://www.r-project.org/Dahan, D., Jude, B. A., Lamendella, R., Keesing, F. & Perron, G. G. Exposure to arsenic alters the Microbiome of larval zebrafish. Front. Microbiol. 9, 1–12 (2018).Article 

    Google Scholar 
    Edgar, R. C. Sequence analysis updating the 97% identity threshold for 16S ribosomal RNA OTUs. Bioinformatics 34, 2371–2375 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wright, E. S. Using DECIPHER v2.0 to analyze big biological sequence data in R. R J. 8, 352–359 (2016).Article 

    Google Scholar 
    Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    McMurdie, P. J. & Holmes, S. Phyloseq: An R package for reproducible interactive analysis and graphics of Microbiome census data. PLoS One. 8, e61217 (2013).Article 
    CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    Karstens, L. et al. Controlling for contaminants in low-biomass 16S rRNA gene sequencing experiments. mSystems 4, e00290 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mikryukov, V. & Mahé, F. vmikk/metagMisc. [Online]. Preprint at (2025). https://doi.org/10.5281/zenodo.597622Caruso, V., Song, X., Asquith, M. & Karstens, L. Performance of Microbiome sequence inference methods in environments with varying biomass. mSystems 4, e00163–e00118 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A. & Callahan, B. J. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome 6, 226 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Knights, D. et al. Bayesian community-wide culture-independent microbial source tracking. Nat. Methods. 8, 761–763 (2013).Article 

    Google Scholar 
    Lahti, L., Shetty, S., Blake, T. & Salajarvi, J. Tools for microbiome analysis in R. (2017). http://microbiome.github.com/microbiomeOksanen, J. et al. vegan: Community ecology package. Preprint at (2019). http://cran.r-project.org/package=veganMurray, D. Tableau your data! Fast and Easy Visual Analysis with Tableau Software (Wiley, 2013).Mauri, M., Elli, T., Caviglia, G., Uboldi, G. & Azzi, M. RAWGraphs: A visualisation platform to create open outputs. ACM Int. Conf. Proc. Series Part F131371 (2017).Download referencesAcknowledgementsWe thank Manuel Antonio Zamarrón Nunez for his assistance during navigation and Ana Sofia Merino, Aurora Paniagua, Madeleine Gauthier, Daniel Valdivia and Ricardo Mirsha Mata Cruz for their help during fieldwork.FundingCADS was funded by a CONACYT PhD Studentship (558253). Fieldwork (sampling and navigation) was funded by The Instituto Politécnico Nacional (SIP20160496 and 2017014), Rufford Foundation 2nd small grant for Nature Conservation (2017), and the Program for the Conservation of Species at Risk (Programa de Conservación de Especies en Riesgo, Comisión Nacional de Áreas Naturales Protegidas). Molecular analysis was partly financed by a Small Grant in Aid of Research from the Society for Marine Mammalogy.Author informationAuthors and AffiliationsUnit for Basic and Applied Microbiology, School of Natural Sciences, Autonomous University of Queretaro, 76010, Queretaro, MexicoCarlos A. Domínguez-Sánchez, Roberto C. Álvarez-Martínez & Karina Acevedo-WhitehouseMingan Island Cetacean Study (MICS), 378 Rue du Bord de la Mer, Longue-Pointe-de-Mingan, QC, G0G 1V0, CanadaCarlos A. Domínguez-SánchezInstituto Politécnico Nacional, Centro Interdisciplinario de Ciencias Marinas, 23096, La Paz, Baja California Sur., MexicoDiane GendronCarr. a Chichimequillas S/N, Ejido Bolaños, 76140, Santiago de Querétaro, Qro, MexicoKarina Acevedo-WhitehouseAuthorsCarlos A. Domínguez-SánchezView author publicationsSearch author on:PubMed Google ScholarDiane GendronView author publicationsSearch author on:PubMed Google ScholarRoberto C. Álvarez-MartínezView author publicationsSearch author on:PubMed Google ScholarKarina Acevedo-WhitehouseView author publicationsSearch author on:PubMed Google ScholarContributionsC.A.D. collected the samples, performed molecular analyses, analyzed the data, and drafted the manuscript. R.C.A. conducted statistical programming for microbiome analysis and helped interpret the results. D.G. conducted fieldwork, collected samples, and co-supervised the research. K.A.W. conceived, designed, and supervised the research. All authors read and commented on the final draft of the manuscript and gave approval for publication.Corresponding authorCorrespondence to
    Karina Acevedo-Whitehouse.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethical approval
    This study complied with the recommendations and methods for approaching blue whales provided by Mexican legislation (NOM-059-SEMARNAT-2010). All procedures were approved by the Bioethics Committee of the Universidad Autónoma de Queretaro (Mexico), and sampling was conducted under permits SGPA/DGVS/00255/16 and SGPA/DGVS/01832/17 issued by the Dirección General de Vida Silvestre to D. Gendron.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleDomínguez-Sánchez, C.A., Gendron, D., Álvarez-Martínez, R.C. et al. Respiratory bacteriome and its predicted functional profiles in blue whales (Balaenoptera musculus).
    Sci Rep 15, 44434 (2025). https://doi.org/10.1038/s41598-025-28025-2Download citationReceived: 03 February 2025Accepted: 07 November 2025Published: 24 December 2025Version of record: 24 December 2025DOI: https://doi.org/10.1038/s41598-025-28025-2Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    Keywords16S rRNABalaenoptera musculusBlue whaleMicrobiomeBacteriomeFunctional profiling More

  • in

    Identifying South African marine protected areas at risk from marine heatwaves and cold-spells

    AbstractMarine heatwaves (MHWs) and marine cold-spells (MCSs) can negatively impact biodiversity as species distributions are largely governed by temperature linked to physiological tolerances. These extremes have not been considered in South Africa’s Marine Protected Area (MPA) network design, so understanding frequency and severity of extreme thermal events will be important for assessing their impact. This study characterises MHWs and MCSs in MPAs across the six South African marine ecoregions, using a novel index to compare thermal event severity. Thermal events declined in duration and intensity from west to east, with the least severe events recorded in the Delagoa ecoregion. Walker Bay MPA was identified as most at risk due to the combined impact of MHWs and MCSs. These thermal events may threaten the ability of the MPA to meet its conservation objective as a cetacean sanctuary. If past trends in MHW frequency and cumulative intensity persist, the majority of South African MPAs could experience more severe heatwaves in the future. Our approach will help prioritise sites for in situ monitoring of water temperature and studies of the impact of extreme thermal events, as well as identifying areas for expanding refugia and conservation corridors, supporting adaptive management into the future.

    Data availability

    Data is freely available on Zenodo: Courtaillac et al. (2024) Identifying South African Marine Protected Areas at risk from marine heatwaves and cold spells [Data set]. Zenodo. DOI: 10.5281/zenodo.14900260.
    ReferencesBryndum-Buchholz, A. et al. Twenty‐first‐century climate change impacts on marine animal biomass and ecosystem structure across ocean basins. Glob. Change Biol. 25, 459–472 (2019).
    Google Scholar 
    Jentsch, A., Kreyling, J. & Beierkuhnlein, C. A new generation of climate-change experiments: Events, not trends. Front. Ecol. Environ. 5, 365–374 (2007).
    Google Scholar 
    Smale, D. A. et al. Marine heatwaves threaten global biodiversity and the provision of ecosystem services. Nat. Clim. Change. 9, 306–312 (2019).
    Google Scholar 
    Feng, M. et al. Multi-year marine cold-spells off the West Coast of Australia and effects on fisheries. J. Mar. Syst. 214, 103473 (2021).
    Google Scholar 
    Steele, J. H., Brink, K. H. & Scott, B. E. Comparison of marine and terrestrial ecosystems: suggestions of an evolutionary perspective influenced by environmental variation. ICES J. Mar. Sci. 76, 50–59 (2019).
    Google Scholar 
    Hutchings, L. et al. The Benguela Current: an ecosystem of four components. Prog. Oceanogr. 83, 15–32 (2009).
    Google Scholar 
    Lutjeharms, J. R. E. Three decades of research on the greater agulhas current. Ocean Sci. 3, 129–147 (2007).
    Google Scholar 
    Hobday, A. J. et al. A hierarchical approach to defining marine heatwaves. Prog. Oceanogr. 141, 227–238 (2016).
    Google Scholar 
    Schlegel, R. W., Oliver, E. C. J., Wernberg, T. & Smit, A. J. Nearshore and offshore co-occurrence of marine heatwaves and cold-spells. Prog. Oceanogr. 151, 189–205 (2017).
    Google Scholar 
    Oliver, E. C. J. et al. Projected marine heatwaves in the 21st century and the potential for ecological impact. Front. Mar. Sci. 6, 1–12 (2019).
    Google Scholar 
    Coleman, M. A., Minne, A. J. P., Vranken, S. & Wernberg, T. Genetic tropicalisation following a marine heatwave. Sci. Rep. 10, 1–11 (2020).
    Google Scholar 
    Wernberg, T. et al. An extreme climatic event alters marine ecosystem structure in a global biodiversity hotspot. Nat. Clim. Change. 3, 78–82 (2013).
    Google Scholar 
    Rogers-Bennett, L. & Catton, C. A. Marine heat wave and multiple stressors tip bull Kelp forest to sea urchin barrens. Sci. Rep. 9, 1–9 (2019).
    Google Scholar 
    Thomson, J. A. et al. Extreme temperatures, foundation species, and abrupt ecosystem change: an example from an iconic seagrass ecosystem. Glob. Change Biol. 21, 1463–1474 (2015).
    Google Scholar 
    Nohaïc, M. L. et al. Marine heatwave causes unprecedented regional mass bleaching of thermally resistant corals in Northwestern Australia. Sci. Rep. 7, 1–11 (2017).
    Google Scholar 
    Tuckett, C. A., de Bettignies, T., Fromont, J. & Wernberg, T. Expansion of corals on temperate reefs: direct and indirect effects of marine heatwaves. Coral Reefs. 36, 947–956 (2017).
    Google Scholar 
    Schlegel, R. W., Darmaraki, S., Benthuysen, J. A., Filbee-Dexter, K. & Oliver, E. C. J. Marine cold-spells. Prog. Oceanogr. 198, 102684 (2021).
    Google Scholar 
    Lubitz, N. et al. Climate change-driven cooling can kill marine megafauna at their distributional limits. Nat. Clim. Chang. 14, 526–535 (2024).
    Google Scholar 
    Canning-Clode, J., Fowler, A. E., Byers, J. E., Carlton, J. T. & Ruiz, G. M. Caribbean creep’ chills out: climate change and marine invasive species. PLoS ONE. 6, e29657 (2011).
    Google Scholar 
    Pinsky, M. L., Selden, R. L. & Kitchel, Z. J. Climate-driven shifts in marine species ranges: scaling from organisms to communities. Annual Rev. Mar. Sci. 12, 153–179 (2020).
    Google Scholar 
    Leriorato, J. C. & Nakamura, Y. Unpredictable extreme cold events: a threat to range-shifting tropical reef fishes in temperate waters. Mar. Biol. 166, 1–10 (2019).
    Google Scholar 
    Sink, K. J. et al. Chapter 3: Marine ecosystem classification and mapping. In National Biodiversity Assessment 2018 Technical Report Volume 4: Marine Realm (eds. Sink, K. J. et al.) 49–108 (South African National Biodiversity Institute, 2019).Hopkins, C. R., Bailey, D. M. & Potts, T. Perceptions of practitioners: managing marine protected areas for climate change resilience. Ocean. Coastal. Manage. 128, 18–28 (2016).
    Google Scholar 
    Fleischman, M. J., Bolton, J. J. & Rothman, M. D. Macrocystis (Laminariales, Phaeophyceae) in South africa: distribution, morphology, and potential susceptibility to warming events. J. Appl. Phycol. 32, 2447–2457 (2020).
    Google Scholar 
    Ackland, S. J., Martin, N. & Robinson, T. B. A temperate Whelk overcomes winter cold spells but not summer heatwaves. Mar. Biol. Res. 21, 198–205 (2025).
    Google Scholar 
    Bates, A. E. et al. Climate resilience in marine protected areas and the ‘Protection paradox’. Biol. Conserv. 236, 305–314 (2019).
    Google Scholar 
    Roberts, C. M. et al. Marine reserves can mitigate and promote adaptation to climate change. Proc. Natl. Acad. Sci. U.S.A. 114, 6167–6175 (2017).
    Google Scholar 
    Bruno, J. F. et al. Climate change threatens the world’s marine protected areas. Nat. Clim. Change. 8, 499–503 (2018).
    Google Scholar 
    Ainsworth, T. D., Hurd, C. L., Gates, R. D. & Boyd, P. W. How do we overcome abrupt degradation of marine ecosystems and meet the challenge of heat waves and climate extremes? Glob. Change Biol. 26, 343–354 (2020).
    Google Scholar 
    Ziegler, S. L. et al. Marine protected areas, marine heatwaves, and the resilience of nearshore fish communities. Sci. Rep. 13, 1405 (2023).
    Google Scholar 
    Arafeh-Dalmau, N. et al. Integrating climate adaptation and transboundary management: guidelines for designing climate-smart marine protected areas. One Earth. 6, 1523–1541 (2023).
    Google Scholar 
    Trisos, C. et al. Africa. In Climate Change 2022 – Impacts, Adaptation and Vulnerability: Working Group II Contribution To the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (ed. Pörtner, O.) (Cambridge University Press, 2023). https://doi.org/10.1017/9781009325844.García-Reyes, M. et al. Most Eastern boundary upwelling regions represent thermal refugia in the age of climate change. Front. Mar. Sci. 10, 1158472 (2023).
    Google Scholar 
    Abrahams, A., Schlegel, R. W. & Smit, A. J. Variation and change of upwelling dynamics detected in the world’s Eastern boundary upwelling systems. Front. Mar. Sci. 8, 626411 (2021).
    Google Scholar 
    Field, J. G., Griffiths, C. L., Linley, E. A., Carter, R. A. & Zoutendyk, P. Upwelling in a nearshore marine ecosystem and its biological implications. Estuar. Coast. Mar. Sci. 11, 133–150 (1980).
    Google Scholar 
    Sherley, R. B. et al. Bottom-up effects of a no-take zone on endangered Penguin demographics. Biol. Lett. 11, 1452 (2015).Blamey, L. K. et al. Ecosystem change in the Southern Benguela and the underlying processes. J. Mar. Syst. 144, 9–29 (2015).
    Google Scholar 
    Kock, A. A. et al. Table Mountain National Park Marine Protected Area—State of Knowledge Report. Internal Report 01/2022. Pretoria: Scientific Services, South African National Parks. 134pp (2022).Steyn, C. & Robinson, T. First record of the marine bryozoan Membranipora membranacea (Linnaeus, 1767) in South Africa. BioInvasion Records (2025).Scheibling, R. & Gagnon, P. Temperature-mediated outbreak dynamics of the invasive bryozoan Membranipora membranacea in Nova Scotian Kelp beds. Mar. Ecol. Prog Ser. 390, 1–13 (2009).
    Google Scholar 
    Griffiths, C. L. & Robinson, T. B. Use and usefulness of measures of marine endemicity in South Africa. South Afr. J. Sci. 112, 1–7 (2016).
    Google Scholar 
    Hutchings, L. et al. Spawning on the edge: spawning grounds and nursery areas around the Southern African coastline. Mar. Freshw. Res. 53, 307 (2002).
    Google Scholar 
    Kirkman, S. P. et al. Evaluating the evidence for ecological effectiveness of South africa’s marine protected areas. Afr. J. Mar. Sci. 43, 389–412 (2021).
    Google Scholar 
    Gissi, E. et al. A review of the combined effects of climate change and other local human stressors on the marine environment. Sci. Total Environ. 755, 142564 (2021).
    Google Scholar 
    Sink, K. J. et al. Chapter 8: Ecosystem protection level. In South African National Biodiversity Assessment 2018 Technical Report Volume 4: Marine Realm (eds Sink, K. J. et al.) 283–326 (South African National Biodiversity Institute, 2019).Adams, R. J. & Kowalski, P. State of Marine Protected Area Management Effectiveness in South Africa. World Wide Fund for Nature, Cape Town South Africa. 140pp (2021).Dransfield, A. et al. Where the whales are: using habitat modeling to support changes in shipping regulations within national marine sanctuaries in central California. Endanger. Species Res. 26, 39–57 (2014).
    Google Scholar 
    Gabriele, C. M. et al. Sharp decline in humpback Whale (Megaptera novaeangliae) survival and reproductive success in southeastern Alaska during and after the 2014–2016 Northeast Pacific marine heatwave. Mammalian Biol. 102, 1113–1131 (2022).
    Google Scholar 
    Lutjeharms, J. R. E., Cooper, J. & Roberts, M. Upwelling at the inshore edge of the Agulhas Current. Cont. Shelf Res. 20, 737–761 (2000).
    Google Scholar 
    Duncan, M., James, N., Bates, A., Goschen, W. & Potts, W. Localised intermittent upwelling intensity has increased along South Africa’s South Coast due to El Niño –Southern Oscillation phase state. Afr. J. Mar. Sci. 41, 325–330 (2019).
    Google Scholar 
    De Oliveira Soares, M. Marginal reef paradox: a possible refuge from environmental changes? Ocean. Coastal. Manage. 185, 105063 (2020).
    Google Scholar 
    Wilson, K. L., Tittensor, D. P., Worm, B. & Lotze, H. K. Incorporating climate change adaptation into marine protected area planning. Glob. Change Biol. 26, 3251–3267 (2020).
    Google Scholar 
    Celliers, L. & Schleyer, M. H. Coral bleaching on high-latitude marginal reefs at Sodwana Bay, South Africa. Mar. Pollut. Bull. 44, 1380–1387 (2002).
    Google Scholar 
    Molewa, B. E. E. Declaration of Amathole Marine Protected Area in the Amathole Region Under Section 43 of the Marine Living Resources Act, 1998 (Act 18 of 1998. Government Gazette (2011).Smit, A. J. et al. A Coastal seawater temperature dataset for biogeographical studies: large biases between in situ and remotely-sensed data sets around the coast of South Africa. PLoS ONE. 8, e81944 (2013).
    Google Scholar 
    Straub, S. C. et al. Persistence of seaweed forests in the anthropocene will depend on warming and marine heatwave profiles. J. Phycol. 58, 22–35 (2022).
    Google Scholar 
    Doxa, A. et al. 4D marine conservation networks: combining 3D prioritization of present and future biodiversity with climatic refugia. Glob. Change Biol. 28, 4577–4588 (2022).
    Google Scholar 
    Elzahaby, Y. & Schaeffer, A. Observational insight into the subsurface anomalies of marine heatwaves. Front. Mar. Sci. 6, 745 (2019).
    Google Scholar 
    Griffiths, C. L., Robinson, T. B., Lange, L. & Mead, A. Marine biodiversity in South Africa: an evaluation of current states of knowledge. PLoS ONE. 5, e12008 (2010).
    Google Scholar 
    Schlegel, R. W. & Smit, A. J. Heatwave R: a central algorithm for the detection of heatwaves and cold-spells. J. Open. Source Softw. 3, 821 (2018).Reynolds, R. W. et al. Daily high-resolution-blended analyses for sea surface temperature. J. Clim. 20, 5473–5496 (2007).
    Google Scholar 
    Banzon, V., Smith, T. M., Chin, T. M., Liu, C. & Hankins, W. A long-term record of blended satellite and in situ sea-surface temperature for climate monitoring, modeling and environmental studies. Earth Syst. Sci. Data. 8, 165–176 (2016).
    Google Scholar 
    Huang, B. et al. Improvements of the daily optimum interpolation sea surface temperature (DOISST) version 2.1. J. Clim. 34, 2923–2939 (2021).
    Google Scholar 
    R Core Team. R. A Language and environment for statistical computing. R foundation for statistical computing R foundation for statistical computing (2023).Pinheiro, J. & Bates, D. & R Core Team. Nlme: Linear and Nonlinear Mixed Effects Models, vol. 3 1–166 (2023).Wickham, H. et al. Welcome to the tidyverse. JOSS 4, 1686 (2019).
    Google Scholar 
    Wickham, H. & Henry, L. Functional Programming Tools. R package version 1.2.0, https://purrr.tidyverse.org/ (2023).Cochrane, D. & Orcutt, G. H. Application of least squares regression to relationships containing auto-correlated error terms. J. Am. Stat. Assoc. 44, 32–61 (1949).
    Google Scholar 
    Stefano, S., Quartagno, M., Tamburini, M. & Robinson, D. Orcutt: estimate procedure in case of first order autocorrelation. R package version 2.3 https://CRAN.R-project.org/package=orcutt (2018).Download referencesAcknowledgementsThe authors gratefully acknowledge the Marine Heatwaves International Working Group and tools that they make available in support of research into extreme thermal events. Sarah Ackland is thanked for help with the formatting of the map. Three anonymous reviewers are thanked for their insightful and in-depth comments that helped to strengthen this work.Author informationAuthors and AffiliationsDepartment of Botany and Zoology, Stellenbosch University, Stellenbosch, Western Cape, South AfricaKira-Lee Courtaillac & Tamara B. RobinsonCape Research Centre, South African National Parks, Cape Town, Western Cape, South AfricaNicola J. van WilgenCentre for Invasion Biology, Stellenbosch University, Stellenbosch, Western Cape, South AfricaNicola J. van Wilgen & Tamara B. RobinsonAuthorsKira-Lee CourtaillacView author publicationsSearch author on:PubMed Google ScholarNicola J. van WilgenView author publicationsSearch author on:PubMed Google ScholarTamara B. RobinsonView author publicationsSearch author on:PubMed Google ScholarContributionsKira-Lee Courtaillac extracted the data, analysed the data, prepared the figures and tables, authored and reviewed drafts of the article and approved the final draft. Nicola van Wilgen conceptualised the study, guided analyses, reviewed drafts of the article and approved the final draft. Tamara Robinson conceptualised the study, analysed the data, prepared tables and figures, authored and reviewed drafts of the article and approved the final draft.Corresponding authorCorrespondence to
    Tamara B. Robinson.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleCourtaillac, KL., van Wilgen, N.J. & Robinson, T.B. Identifying South African marine protected areas at risk from marine heatwaves and cold-spells.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-32725-0Download citationReceived: 21 February 2025Accepted: 11 December 2025Published: 24 December 2025DOI: https://doi.org/10.1038/s41598-025-32725-0Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsAgulhas currentBenguela currentClimate changeExtreme thermal eventsMarine conservation planning More

  • in

    Roles of micro/nanoplastics in the spread of antimicrobial resistance through conjugative gene transfer

    AbstractThe role of micro/nanoplastics (M/NPs) in the dissemination of antimicrobial resistance (AMR) remains insufficiently understood. Here, we examine how polystyrene (PS) M/NPs of varying sizes and concentrations affect AMR gene (ARG) transfer in model systems with gram-negative (Escherichia coli) and gram-positive (Enterococcus faecalis) donors. In these systems, the ARG transfer frequency is higher for intrageneric pairs than for intergeneric pairs. The 20- and 120-nm-sized PS broadly facilitate conjugation, whereas the 1-μm-sized PS selectively promotes ARG transfer to E. coli recipients, in addition to altering the expression of conjugation- and pili-associated genes. Notably, an environmentally relevant (0.1 mg/L) concentration of PS M/NPs facilitates AMR transfer in the tested systems, which correlates with increased reactive oxygen species levels, ATP levels, and cell membrane permeability in both donors and recipients. Collectively, our findings underscore the role of M/NPs in facilitating AMR spread in specific bacterial systems, providing valuable insights for understanding their potential ecological risk in water environments.

    Data availability

    All data generated or analysed during this study are included in this published article, its Supplementary Information and the accompanying Source Data file. All RNA sequencing data have been deposited in the NCBI Gene Expression Omnibus under accession codes GSE248909 and GSE297944. Source data are provided with this paper.
    ReferencesDawson, A. L. et al. Turning microplastics into nanoplastics through digestive fragmentation by Antarctic krill. Nat. Commun. 9, 1001 (2018).Koelmans, A. A. et al. Microplastics in freshwaters and drinking water: Critical review and assessment of data quality. Water Res 155, 410–422 (2019).
    Google Scholar 
    Rochman, C. M. Microplastics research-from sink to source. Science 360, 28–29 (2018).
    Google Scholar 
    Wu, N. Tracing microplastic footprints through the plastisphere. Nat. Rev. Earth Environ. 3, 498–498 (2022).
    Google Scholar 
    Materic, D. et al. Micro- and nanoplastics in alpine snow: a new method for chemical identification and (semi)quantification in the nanogram range. Environ. Sci. Technol. 54, 2353–2359 (2020).
    Google Scholar 
    Seeley, M. E., Song, B., Passie, R. & Hale, R. C. Microplastics affect sedimentary microbial communities and nitrogen cycling. Nat. Commun. 11, 2372 (2020).
    Google Scholar 
    Yan, Z. et al. Analysis of microplastics in human feces reveals a correlation between fecal microplastics and inflammatory bowel disease status. Environ. Sci. Technol. 56, 414–421 (2022).
    Google Scholar 
    Koelmans, A. A. et al. Risk assessment of microplastic particles. Nat. Rev. Mater. 7, 138–152 (2022).
    Google Scholar 
    Allen, D. et al. Microplastics and nanoplastics in the marine-atmosphere environment. Nat. Rev. Earth Environ. 3, 393–405 (2022).
    Google Scholar 
    Kumar, M. et al. Current research trends on micro- and nano-plastics as an emerging threat to global environment: a review. J. Hazard Mater. 409, 124967 (2021).
    Google Scholar 
    MacLeo, M., Arp, H. P. H., Tekman, M. B. & Jahnke, A. The global threat from plastic pollution. Science 373, 61–65 (2021).
    Google Scholar 
    Andersson, D. I. & Hughes, D. Microbiological effects of sublethal levels of antibiotics. Nat. Rev. Microbiol. 12, 465–478 (2014).
    Google Scholar 
    Zainab, S. M., Junaid, M., Xu, N. & Malik, R. N. Antibiotics and antibiotic resistant genes (ARGs) in groundwater: a global review on dissemination, sources, interactions, environmental and human health risks. Water Res. 187, 116455 (2020).
    Google Scholar 
    Ellabaan, M. M. H., Munck, C., Porse, A., Imamovic, L. & Sommer, M. O. A. Forecasting the dissemination of antibiotic resistance genes across bacterial genomes. Nat. Commun. 12, 2435 (2021).
    Google Scholar 
    Jiang, X. et al. Dissemination of antibiotic resistance genes from antibiotic producers to pathogens. Nat. Commun. 8, 15784 (2017).
    Google Scholar 
    May, M. Tomorrow’s biggest microbial threats. Nat. Med. 27, 358–359 (2021).
    Google Scholar 
    Zhang, Z. Y. et al. Assessment of global health risk of antibiotic resistance genes. Nat. Commun. 13, 1553 (2022).Lu, J. & Guo, J. H. Disinfection spreads antimicrobial resistance. Science 371, 474–474 (2021).
    Google Scholar 
    Wang, Y. et al. Non-antibiotic pharmaceuticals promote the transmission of multidrug resistance plasmids through intra- and intergenera conjugation. ISME J. 15, 2493–2508 (2021).
    Google Scholar 
    Lu, J. et al. Both silver ions and silver nanoparticles facilitate the horizontal transfer of plasmid-mediated antibiotic resistance genes. Water Res. 169, 115229 (2020).Zhang, Y., Gu, A. Z., He, M., Li, D. & Chen, J. Subinhibitory concentrations of disinfectants promote the horizontal transfer of multidrug resistance genes within and across genera. Environ. Sci. Technol. 51, 570–580 (2017).
    Google Scholar 
    Shi, X., Xia, Y., Wei, W. & Ni, B. J. Accelerated spread of antibiotic resistance genes (ARGs) induced by non-antibiotic conditions: Roles and mechanisms. Water Res. 224, 119060 (2022).
    Google Scholar 
    Redondo-Hasselerharm, P. E. Gort, G. Peeters, E. & Koelmans, A. A. Nano- and microplastics affect the composition of freshwater benthic communities in the long term. Sci. Adv. 6 eaay4054 (2020).Li, B. et al. Metagenomic and network analysis reveal wide distribution and co-occurrence of environmental antibiotic resistance genes. ISME J. 9, 2490–2502 (2015).
    Google Scholar 
    Hu, X., Waigi, M. G., Yang, B. & Gao, Y. Impact of plastic particles on the horizontal transfer of antibiotic resistance genes to bacterium: dependent on particle sizes and antibiotic resistance gene vector replication capacities. Environ. Sci. Technol. 56, 14948–14959 (2022).
    Google Scholar 
    Brito, I. L. Examining horizontal gene transfer in microbial communities. Nat. Rev. Microbiol. 19, 442–453 (2021).
    Google Scholar 
    Liu, L., Zhang, Q.-H. & Li, R.-T. In situ and individual-based analysis of the influence of polystyrene microplastics on Escherichia coli conjugative gene transfer at the single-cell level. Environ. Sci. Technol. 57, 15936–15944 (2023).
    Google Scholar 
    Xu, Y., Ou, Q., Jiao, M., Liu, G. & van der Hoek, J. P. Identification and quantification of nanoplastics in surface water and groundwater by pyrolysis gas chromatography–mass spectrometry. Environ. Sci. Technol. 56, 4988–4997 (2022).
    Google Scholar 
    Della Torre, C. et al. Accumulation and embryotoxicity of polystyrene nanoparticles at early stage of development of sea urchin embryos Paracentrotus lividus. Environ. Sci. Technol. 48, 12302–12311 (2014).
    Google Scholar 
    Ivleva, N. P. Chemical analysis of microplastics and nanoplastics: challenges, advanced methods, and perspectives. Chem. Rev. 121, 11886–11936 (2021).
    Google Scholar 
    Enfrin, M. et al. Release of hazardous nanoplastic contaminants due to microplastics fragmentation under shear stress forces. J. Hazard. Mater. 384, 121393 (2020).Redondo-Salvo, S. et al. Pathways for horizontal gene transfer in bacteria revealed by a global map of their plasmids. Nat. Commun. 11, 3602 (2020).
    Google Scholar 
    Rodríguez-Beltrán, J. et al. Beyond horizontal gene transfer: the role of plasmids in bacterial evolution. Nat. Rev. Microbiol. 19, 347–359 (2021).
    Google Scholar 
    Cheng, P. et al. Fe-doped TiO2 nanosheet exposure accelerates the spread of antibiotic resistance genes by promoting plasmid-mediated conjugative transfer. J. Hazard. Mater. 490, 137715 (2025).
    Google Scholar 
    Yu, K. et al. CeO2 Nanoparticles regulate the propagation of antibiotic resistance genes by altering cellular contact and plasmid transfer. Environ. Sci. Technol. 54, 10012–10021 (2020).
    Google Scholar 
    Kohanski, M. A., DePristo, M. A. & Collins, J. J. Sublethal antibiotic treatment leads to multidrug resistance via radical-induced mutagenesis. Mol. Cell 37, 311–320 (2010).
    Google Scholar 
    Huang, S. et al. Metagenomic analysis reveals the responses of microbial communities and nitrogen metabolic pathways to polystyrene micro(nano)plastics in activated sludge systems. Water Res. 241, 120161 (2023).
    Google Scholar 
    Duan, J. J. et al. ROS-mediated photoaging pathways of nano- and micro-plastic particles under UV irradiation. Water Res. 216, 118320 (2022).Kohanski, M. A., Dwyer, D. J. & Collins, J. J. How antibiotics kill bacteria: from targets to networks. Nat. Rev. Microbiol. 8, 423–435 (2010).
    Google Scholar 
    Ho, B. T., Basler, M. & Mekalanos, J. J. Type 6 secretion system-mediated immunity to type 4 secretion system-mediated gene transfer. Science 342, 250–253 (2013).
    Google Scholar 
    DeBenedictis, E. P., Liu, J., & Keten, S. Adhesion mechanisms of curli subunit CsgA to abiotic surfaces. Sci. Adv. 2 (2016).Yu, Z. G., Wang, Y., Lu, J., Bond, P. L. & Guo, J. H. Nonnutritive sweeteners can promote the dissemination of antibiotic resistance through conjugative gene transfer. ISME J. 15, 2117–2130 (2021).
    Google Scholar 
    Maier, B. & Wong, G. C. L. How bacteria use type IV pili machinery on surfaces. Trends Microbiol. 23, 775–788 (2015).
    Google Scholar 
    Dunny, G. M. & Berntsson, R. P. Enterococcal sex pheromones: evolutionary pathways to complex, two-signal systems. J. Bacteriol. 198, 1556–1562 (2016).
    Google Scholar 
    Kohler, V., Keller, W. & Grohmann, E. Enterococcus adhesin PrgB facilitates type IV secretion by condensation of extracellular DNA. Mol. Microbiol. 109, 263–267 (2018).
    Google Scholar 
    Thomas, C. M. & Nielsen, K. M. Mechanisms of, and barriers to, horizontal gene transfer between bacteria. Nat. Rev. Microbiol. 3, 711–721 (2005).
    Google Scholar 
    Alimi, O. S., Budarz, J. F., Hernandez, L. M. & Tufenkji, N. Microplastics and nanoplastics in aquatic environments: aggregation, deposition, and enhanced contaminant transport. Environ. Sci. Technol. 52, 1704–1724 (2018).
    Google Scholar 
    Zha, Y. Y. et al. Size-dependent enhancement on conjugative transfer of antibiotic resistance genes by micro/nanoplastics. J. Hazard. Mater. 431, 128561 (2022).von Wintersdorff, C. J. et al. Dissemination of antimicrobial resistance in microbial ecosystems through horizontal gene transfer. Front. Microbiol. 7, 173 (2016).
    Google Scholar 
    Wang, X. X. et al. A neglected risk of nanoplastics as revealed by the promoted transformation of plasmid-borne ampicillin resistance gene by. Environ. Microbiol. 24, 4946–4959 (2022).
    Google Scholar 
    Besseling, E., Redondo-Hasselerharm, P., Foekema, E. M. & Koelmans, A. A. Quantifying ecological risks of aquatic micro- and nanoplastic. Crit. Rev. Environ. Sci. Technol. 49, 32–80 (2019).
    Google Scholar 
    Zhao, Y. et al. Size-dependent promotion of micro(nano)plastics on the horizontal gene transfer of antibiotic resistance genes in constructed wetlands. Water Res. 244, 120520 (2023).
    Google Scholar 
    Xu, Y., Li, H., Ding, Y., Zhang, D. & Liu, W. How nanoscale plastics facilitate the evolution of antibiotic resistance?. J. Hazard. Mater. 480, 136157 (2024).
    Google Scholar 
    Liu, L., Sun, Y., Du, S., Li, Y. & Wang, J. Nanoplastics promote the dissemination of antibiotic resistance genes and diversify their bacterial hosts in soil. Eco-Environ. Health 3, 1–10 (2024).
    Google Scholar 
    Bottery, M. J., Wood, A. J. & Brockhurst, M. A. Temporal dynamics of bacteria-plasmid coevolution under antibiotic selection. ISME J. 13, 559–562 (2019).
    Google Scholar 
    Liu, X. et al. Nano- and microplastics aided by extracellular polymeric substances facilitate the conjugative transfer of antibiotic resistance genes in bacteria. ACS ES&T Water 2, 2528–2537 (2022).
    Google Scholar 
    Yu, X. et al. Microplastics exacerbate co-occurrence and horizontal transfer of antibiotic resistance genes. J. Hazard Mater. 451, 131130 (2023).
    Google Scholar 
    Chen, M.-M., Zhang, Y.-Q., Cheng, L.-C., Zhao, F.-J. & Wang, P. Photoaged nanoplastics with multienzyme-like activities significantly shape the horizontal transfer of antibiotic resistance genes. J. Hazard. Mater. 475, 134884 (2024).
    Google Scholar 
    Zhou, C. S. et al. Micro) nanoplastics promote the risk of antibiotic resistance gene propagation in biological phosphorus removal system. J. Hazard. Mater. 431, 128547 (2022).
    Google Scholar 
    Wu, Y. et al. Biochar effectively inhibits the horizontal transfer of antibiotic resistance genes via restraining the energy supply for conjugative plasmid transfer. Environ. Sci. Technol. 423, 127150 (2022).
    Google Scholar 
    Klümper, U. et al. Metal stressors consistently modulate bacterial conjugal plasmid uptake potential in a phylogenetically conserved manner. ISME J. 11, 152–165 (2017).
    Google Scholar 
    Gigault, J. et al. Nanoplastics are neither microplastics nor engineered nanoparticles. Nat. Nanotechnol. 16, 501–507 (2021).
    Google Scholar 
    Huang, H., Liao, J., Zheng, X., Chen, Y. & Ren, H. Low-level free nitrous acid efficiently inhibits the conjugative transfer of antibiotic resistance by altering intracellular ions and disabling transfer apparatus. Water Res. 158, 383–391 (2019).
    Google Scholar 
    Maier, L. et al. Extensive impact of non-antibiotic drugs on human gut bacteria. Nature 555, 623–628 (2018).
    Google Scholar 
    Mitrano, D. M. & Wohlleben, W. Microplastic regulation should be more precise to incentivize both innovation and environmental safety. Nat. Commun. 11, 5324 (2020).
    Google Scholar 
    Yang, Y. et al. Bisphenols promote the pheromone-responsive plasmid-mediated conjugative transfer of antibiotic resistance genes in Enterococcus faecalis. Environ. Sci. Technol. 56, 17653–17662 (2022).
    Google Scholar 
    Fang, F. C. Antibiotic and ROS linkage questioned. Nat. Biotechnol. 31, 415–416 (2013).
    Google Scholar 
    Download referencesAcknowledgementsThis work was supported by the National Natural Science Foundation of China (grant nos. 52321005 awarded to A.W., 52293441 awarded to S.-H.G., 52293443 awarded to A.W., and 52070060 awarded to S.-H.G.), the Natural Science Foundation of Guangdong Basic and Applied Basic Research Foundation (Grant No. 2024A1515010085 awarded to S.-H.G.), the Shenzhen Overseas High-level Talents Research Startup Program (No. 20200518750C awarded to S.-H.G.), the Shenzhen Overseas High-Level Talent Innovation and Entrepreneurship Special Fund (No. KQTD20190929172630447 awarded to S.-H.G.) and Shenzhen Science and Technology Program (Nos. GXWD20231127195344001 awarded to A.W. and S.-H.G. and JCYJ20241202123735045 awarded to S.-H.G.), and the State Key Laboratory of Urban-rural Water Resource and Environment (Harbin Institute of Technology) (No.2025TS39 awarded to S.-H.G.). We would like to thank Prof. Zhigang Qiu from the Tianjin Institute of Environmental and Operational Medicine for donating the bacterial strains E. faecalis OG1RF and E. faecalis OG1RS, and Dr. Casey Huang and Dr. Lyman Tze Kin Ngiam from the Australian Centre for Water and Environmental Biotechnology for proofreading the paper. Figure 5 was designed, composed, and edited using BioRender (Kang, Y. (2025) https://BioRender.com/ofu614f), ChemDraw, and Adobe Illustrator.Author informationAuthor notesThese authors contributed equally: Yuanyuan Kang, Shu-Hong Gao.Authors and AffiliationsState Key Laboratory of Urban-Rural Water Resource and Environment, School of Eco-Environment, Harbin Institute of Technology, Shenzhen, ChinaYuanyuan Kang, Shu-Hong Gao, Yusheng Pan, Tianyao Li, Yiyi Su, Wanying Zhang, Bin Liang & Aijie WangState Key Laboratory of Urban-Rural Water Resource and Environment, School of Environment, Harbin Institute of Technology, Harbin, ChinaRui GaoDepartment of Ocean Science and Engineering, Southern University of Science and Technology, Shenzhen, ChinaLu FanAustralian Centre for Water and Environmental Biotechnology, The University of Queensland, Brisbane, QLD, AustraliaZhigang Yu & Jianhua GuoState Key Laboratory for Ecological Security of Regions and Cities, Institute of Urban Environment,, Chinese Academy of Sciences, Xiamen, ChinaJian-Qiang SuState Key Laboratory of Water Pollution Control and Green Resource Recycling, School of the Environment, Nanjing University, Nanjing, ChinaYi LuoSchool of Environmental Science and Engineering, Tiangong University, Tianjin, ChinaYue WangAuthorsYuanyuan KangView author publicationsSearch author on:PubMed Google ScholarShu-Hong GaoView author publicationsSearch author on:PubMed Google ScholarYusheng PanView author publicationsSearch author on:PubMed Google ScholarRui GaoView author publicationsSearch author on:PubMed Google ScholarTianyao LiView author publicationsSearch author on:PubMed Google ScholarLu FanView author publicationsSearch author on:PubMed Google ScholarYiyi SuView author publicationsSearch author on:PubMed Google ScholarWanying ZhangView author publicationsSearch author on:PubMed Google ScholarZhigang YuView author publicationsSearch author on:PubMed Google ScholarBin LiangView author publicationsSearch author on:PubMed Google ScholarJian-Qiang SuView author publicationsSearch author on:PubMed Google ScholarYi LuoView author publicationsSearch author on:PubMed Google ScholarYue WangView author publicationsSearch author on:PubMed Google ScholarJianhua GuoView author publicationsSearch author on:PubMed Google ScholarAijie WangView author publicationsSearch author on:PubMed Google ScholarContributionsY.K. and S.-H.G. designed the overall experiments for this study. Y.K. performed all analyses, conducted the RP4-relevant conjugation experiments within and across genera, measured ROS levels, and detected changes in cell membrane permeability and ATP generation in RP4-relevant strains. S.-H.G. supervised and managed the project and contributed to the writing and revision of the manuscript. Y.K. and S.-H.G. wrote the full manuscript and illustrated all the figures provided. Y.P., T.L. and R.G. performed the pCF10-relevant conjugation experiments and corresponding measurements of ROS production, changes in cell membrane permeability and ATP generation; Y.S. and W.Z. performed the experiments, analysed the data, and revised the manuscript. J.G. and Z.Y. contributed to the initial planning for this study, provided guidance on the research significance of this study, and contributed to revising the manuscript. L.F. and B.L. assisted in analysing the mechanisms, ecological significance, and potential application scenarios of this study. J.S. and Y.L. provided the donor and recipient strains and provided feedback on the conjugation experiments. Y.W. analysed the transcriptomic data and determined the changes in the expression of related genes under different concentrations and particle size PS treatments. A.W. provided guidance on the concentration and particle size of PS in the study and contributed to revising the manuscript. All the authors provided feedback and discussed the manuscript.Corresponding authorsCorrespondence to
    Shu-Hong Gao, Yue Wang or Aijie Wang.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Lorenzo Brusetti, Raffaella Sabatino, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationReporting SummaryTransparent Peer Review fileSource dataSource dataRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleKang, Y., Gao, SH., Pan, Y. et al. Roles of micro/nanoplastics in the spread of antimicrobial resistance through conjugative gene transfer.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67879-yDownload citationReceived: 26 May 2025Accepted: 11 December 2025Published: 24 December 2025DOI: https://doi.org/10.1038/s41467-025-67879-yShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More

  • in

    Ecosystem carbon use efficiency at global scale from upscaling eddy-covariance data with machine learning and MODIS products

    AbstractCarbon use efficiency (CUE) is a key indicator in coupled biological–abiotic systems that accounts for their capacity of effectively retain carbon, offering insights of ecosystem functioning and the dynamics of carbon cycle. Generally, CUE assessments have been limited to an autotrophic perspective, quantifying plant efficiency while neglecting carbon losses from heterotrophic respiration. This provides an incomplete view of ecosystem carbon retention. To address this critical gap, we offer a global quantification of a more holistic ecosystem-level CUE, that incorporates all respiratory fluxes. This paper proposes a methodology for mapping ecosystem CUE at global scale from in situ data, remote sensing observations and machine learning. This data-driven approach exploits a Gaussian Processes Regression (GPR) model trained with CUE from eddy-covariance towers and concomitant observations from the Moderate resolution Imaging Spectroradiometer (MODIS). The performance of the model shows high correspondence (R2 = 0.84) and low error and bias (RMSE = 0.1, ME = 0.01) regarding in situ data. The execution of the GPR model upscaled CUE and associated uncertainty from tower level to global scale and provided multitemporal global CUE estimates from 2001 to 2023. The GPR model reports a mean global CUE of 0.43 ± 0.08 for this period. A preliminary analysis carried out for different climatic zones and biomes illustrates the increase of mean CUE from tropical (0.36 ± 0.08) to cold (0.55 ± 0.08) zones. The lowest mean CUE is found over evergreen broadleaved forests (0.37 ± 0.04), whereas the largest mean CUE is found over open shrublands (0.53 ± 0.12). Finally, a global trend of (–1.2 ± 0.3) × 10–3 yr− 1 is reported for mean global CUE from 2001 to 2023. The results of this work highlight the dependence of CUE on both climate and biome type, as well as the decreasing carbon sequestration power of vegetation at global scale, which is key to better understand the effects of climate change.

    Similar content being viewed by others

    Global variation in vegetation carbon use efficiency inferred from eddy covariance observations

    Article
    Open access
    19 June 2025

    Emerging multiscale insights on microbial carbon use efficiency in the land carbon cycle

    Article
    Open access
    13 September 2024

    Reduced water loss rather than increased photosynthesis controls CO2-enhanced water-use efficiency

    Article

    11 July 2025

    IntroductionThe carbon use efficiency (CUE) is an important ecological indicator that quantifies the capacity of terrestrial ecosystems to act as sinks for carbon transferred from the atmosphere. It is considered as one of the three key axes that capture most of the variability within ecosystem functions1; the other two axes are the maximum ecosystem productivity and the ecosystem water-use efficiency. The knowledge of the three axes at global scale is crucial to understand and quantify the response of ecosystems to climatic and other environmental changes and to map the overall ecosystem functioning.“Carbon use efficiency is a conceptually simple parameter. The reality of measuring CUE is, however, methodologically challenging.” This was stated by Bradford & Crowther2 in 2013, who point out that CUE requires measurement of both carbon uptake and associated growth. The gross primary productivity (GPP) –which is not measured in situ directly but obtained from eddy covariance (EC) measurements– is frequently used in the literature to estimate the amount of carbon uptake: it quantifies the carbon fixed by plants through photosynthesis. CUE is usually defined for a plant community as the ratio of net primary productivity (NPP) to GPP, being NPP = GPP – Ra, where Ra refers to the autotrophic (plant) respiration. GPP represents the capacity of plants to transform carbon into new biomass. A higher value of CUE translates to greater growth per unit of carbon mass acquired. However, since carbon can be also stored in soils and thus released to atmosphere by both autotrophs and heterotrophs (microbial or heterotrophic respiration, Rh), the “ecosystem carbon use efficiency” is introduced3,4,5 and defined as the ratio of the net ecosystem productivity (NEP = GPP – (Ra + Rh)) to GPP, CUE = NEP/GPP. Under non-steady-state conditions, NEP is retained in the ecosystem as phytomass and/or soil organic carbon5 and, thus, the ecosystem carbon use efficiency represents the potential carbon sink capacity of the ecosystem. According to this, ecosystem CUE values are lower than those of plant communities. From now on, we use CUE to refer to that obtained using the total ecosystem respiration (Ra + Rh) and CUE’ to refer that including only Ra (Fig. 1). The plant-centric view of CUE’ overlooks the critical role of Rh, which can represent a significant pathway for carbon return to the atmosphere. Although Rh is usually smaller than the other carbon fluxes, it can represent a significant percentage of total ecosystem respiration: 23%–32% for forests, 47%–57% for prairies and tundra, and 48%–54% for agricultural systems6; therefore CUE(:<)CUE’, and this must be considered when comparing carbon use efficiency data from different sources. An exclusive focus on CUE’ may therefore provide an incomplete picture of the ecosystem’s overall capacity to retain carbon. To address this conceptual gap, this study aims to provide a global quantification of ecosystem-level CUE. By defining CUE based on NEP, our approach explicitly incorporates both autotrophic and heterotrophic respiration. This holistic metric moves beyond the producer level to assess the carbon use efficiency of the entire ecosystem offering a more comprehensive assessment of how effectively terrestrial ecosystems assimilate and sequester atmospheric carbon.Fig. 1Schematic definition of ecosystem carbon use efficiency (CUE) and plant-use efficiency (CUE’). Box sizes not scaled.Full size imageEcosystem-atmosphere carbon fluxes can show both spatial and temporal variability7,8 due to changes in ecosystem type and composition, weather patterns, and phenology, among others9. Therefore, both spatial and temporal changes in CUE might be expected10. CUE shows less variability than the carbon fluxes used to compute it. While net and gross productivities and ecosystem respiration vary by two orders of magnitude across biomes11, CUE values remain within a narrower range (if we exclude ecosystems with negative NEP). However, this does not imply a single, universal CUE value for any given ecosystem type. Jin et al.12, using EC data, propose a global average CUE of 0.50 ± 0.13, with the greatest values corresponding with croplands and the lowest with mixed forests. Chen & Yu13 proposed an averaged value of CUE’ = 0.537 ± 0.114 across the whole China region, which showed large spatial variations associated to climate factors (mean annual temperature and) precipitation, and to ecosystem types. Particularly, for forest ecosystems –which are the main carbon sinks–, substantial variation in CUE’ has been reported in the literature14 because tree respiration is not a constant fraction of photosynthesis but it depends on forest age and type. The highest CUE’ values are found in young forests and in temperate deciduous forests.Both the processes of photosynthesis and respiration are affected by environmental conditions such as elevated temperatures, water shortage and increased drought stress typical of global warming15. Therefore, a CUE temporal variability is expected12. CUE is usually related to temperature5 –which is the main controlling factor for CUE variations16– and to precipitation and water availability13. For example, drought events that affected southeast Europe during the 2000–2014 period reduced the CUE by 10 to 20% and, as a result, the region shifted from a carbon sink to a carbon source17.Despite all the variations in spatial and temporal scales, terrestrial carbon cycle models have frequently assumed fixed values of CUE’ or CUE, that is, fixed values for the ratio Ra/GPP or (Ra + Rh)/GPP, respectively. However, the literature has sufficiently shown the weakness of this assumption, highlighting that the knowledge of CUE variability would contribute to improve the carbon fluxes estimates and their prediction to understand ecosystem functioning14,16. This highlights the need to map annual CUE at global scale (establishing differences between ecosystem types) and to analyze temporal and spatial variations to detect areas vulnerable to degradation under external disturbances (negative CUE trend), or even areas that can shift from being a carbon sink to a carbon source.In situ data provide insights into CUE patterns; however, their scarcity, driven by vegetation heterogeneity and limited sampling, constrains accurate CUE mapping at the global scale. For example, Desai et al.18 evaluate the variation in NEP in the upper Great Lakes region of the United States of America (USA) and show the challenges of upscaling the fluxes across space and the limits of using data from very tall EC towers as proxies for regional fluxes. These in situ carbon flux measurements are only representative of an area determined by the tower’s footprint area (local scale): it can range from a few hundred meters to a few kilometers depending on tower height, canopy characteristics (heterogeneity), and wind velocity19. The upscaling of these punctual measurements to global scale can be achieved by exploiting machine learning and deep learning approaches such as random forests20, neural networks21 and Gaussian processes (GPs)7,22, among others.In this framework, the main goal of this manuscript is to provide a tool to obtain ecosystem CUE at the global scale through the upscaling of in situ EC CUE exploiting Gaussian Processes Regression (GPR) and remote sensing (RS) observations from the MODerate resolution Imaging Spectroradiometer (MODIS). To our knowledge, no 1-km global-scale and long-term multitemporal validated carbon use efficiency maps adopting the ecosystem carbon use efficiency definition are currently available. The application of this definition in the existing literature has been predominantly confined to site-specific or regional analyses, thereby lacking the spatial coverage and temporal consistency required for a global-scale assessment. Our approach computes global, long-term CUE estimates, accounting for the contribution of total ecosystem respiration, thereby filling an important gap in ecosystem carbon research. The results are evaluated as a function of climatic zones and biome types, considering –in addition to literature data– the CUE’ that is obtained from the MOD17A3HGF product. To highlight the usefulness of the proposed CUE, changes in annual CUE over the 2001–2023 period are analyzed to determine CUE potential for detecting particularly sensitive or vulnerable areas at the global scale.This manuscript is organized into 4 sections. Both in situ and RS data are described in the section “Materials and methods”, which also includes the description of the upscaling approach to obtain global CUE. The section “Results and discussion” contains the results and discussion concerning CUE at global scale including the spatial variability in terms of climate and ecosystem types, and the annual CUE map obtained throughout time series of 22 years to detect trends. Finally, the section “Conclusions” summarizes the main conclusions. Further details of the results are included in Supplementary Information online.Materials and methodsIn situ dataThe CUE upscaling process implies the training and execution of a machine learning model, which is fed by MODIS observations that are used as model inputs whereas CUE is the model output. In this study, in situ CUE was computed using concomitant GPP and NEP from the FLUXNET network (https://fluxnet.org). The in situ GPP and NEP measurements are based on the use of the EC technique, which is the standard method to measure directly (in situ) trace gas fluxes between ecosystems and atmosphere19. In particular, the most recent FLUXNET data product is used, the FLUXNET2015 Dataset (https://fluxnet.org/data/fluxnet2015-dataset/). Detailed descriptions of this dataset, and the method to partition NEP into GPP and ecosystem respiration are found in Pastorello et al.23 and Reichstein et al.24, respectively. The FLUXNET network allowed us to gather daily data from 211 EC flux towers distributed around the world. The towers are located in almost all the representative ecosystems (mainly forests and grasslands) from 77º N to 57º S19. Some of these sites have been collecting data for several decades, allowing the study of ecosystems over time. Supplementary Fig. S1 available online shows the location, dominant ecosystem, and climate type over every flux tower. Finally, daily data were temporally aggregated to match the temporal resolution (8-day) of the MODIS products. It is worth mentioning that the temporal coverage of EC flux towers in the Fluxnet2015 dataset is not uniform: while some sites only provide between one and four years of observations, others cover more than a decade. A quality control was applied to retain only the highest quality samples, selecting those that simultaneously met the highest quality EC tower data and MODIS observations. This filtering process yielded a total of 2912 samples available for model training and evaluation. The temporal distribution of the samples spans different years and months, with a lower proportion during the December-March period (see Supplementary Fig. S2 online). Although this number may seem limited for a global application, it is comparable to datasets used in other studies that have successfully implemented GPR models for global mapping25.Remote sensing observationsThe inputs selected in this study are RS observations of the MODIS sensor, which are related with environmental and physiological variables, and biosphere-atmosphere interactions. In particular, eight MODIS-based predictors from five different products are selected (see Table 1): the MCD18C2 product26 that provides Photosynthetically Active Radiation (PAR); the MOD11A2 product27, which provides day- and night-time land surface temperature (LSTD, LSTN); the MOD16A2 product28, which derives Evapotranspiration (ET) and Potential Evapotranspiration (PET), and allows to compute a water stress factor (Cws) as Cws = ET/PET; the MCD15A2H product29 that provides the Leaf Area Index (LAI); and the kernel version of the normalized difference vegetation index (kNDVI)30 computed from the MCD43A4 product31. All these products were downloaded at 1-km spatial resolution using Google Earth Engine (GEE) over the locations of the EC towers from 2001 to 2023. As previously mentioned, every product’s quality flag was used to filter only best quality pixels. Eventually, an annual CUE’, computed as CUE’ = NPP/GPP from the MOD17A3HGF product32 has been also downloaded from GEE to be compared with the CUE retrieved by the GPR.Table 1 Variables selected or computed from MODIS products.Full size tableGaussian processes regressionGPs are Bayesian tools for discriminative machine learning33. When applied to regression, they are known as Gaussian process regression, which provides a probabilistic approach to nonparametric kernel-based regression methods. GPR assumes a Gaussian process prior governs the set of possible unobserved (latent) functions. Both the observations and the likelihood yield posterior probabilistic estimates. The standard regression model expresses the output as the sum of an unknown latent function f(x) of the inputs and independent Gaussian noise of the form (:mathcal{N}(0,{sigma:}_{n}^{2})). GPR adopts a Bayesian, non-parametric framework by placing a zero-mean Gaussian process prior (:GPleft(0,{k}_{{uptheta:}}right)) on the latent function. Here, (:{k}_{{uptheta:}}) is a covariance function parameterized by hyperparameters (:varvec{theta:}). Given this prior, samples of f(x) evaluated at input locations (:mathbf{X}=left{{mathbf{x}}^{left(1right)},{mathbf{x}}^{left(2right)},dots:,{mathbf{x}}^{left(Nright)}right}), where N is the number of samples, each one having P features, follow a joint multivariate Gaussian distribution with zero mean and covariance matrix K with (:{left[mathbf{K}right]}_{ij}=:{k}_{{uptheta:}}:({mathbf{x}}^{left(iright)},{mathbf{x}}^{left(jright)}).).For a test input (:{mathbf{x}}_{mathbf{*}}) associated to a scalar output (:{text{y}}_{text{*}}), the GPR induces a prior distribution between the observations y and (:{mathbf{x}}_{mathbf{*}}) with the corresponding output (:{text{y}}_{mathbf{*}}). Given a dataset (:mathcal{D}equiv:{mathbf{X},y}), the posterior distribution over a, unknown output (:{text{y}}_{text{*}}) can be derived analytically as:$$:pleft({y}_{*}left|{mathbf{x}}_{mathbf{*}},mathcal{D}right.right)=mathcal{N}left({y}_{*}left|{mu:}_{GPR*},{sigma:}_{GPR*}^{2}right.right)$$
    (1)
    $$:{mu:}_{GPR*}={mathbf{k}}_{text{*}}^{{top:}}{(mathbf{K}+{sigma:}_{n}^{2}{mathbf{I}}_{text{n}})}^{-1}mathbf{y}={mathbf{k}}_{text{*}}^{{top:}}alpha:$$
    (2)
    $$:{sigma:}_{GPR*}^{2}={sigma:}_{n}^{2}+{k}_{**}{-mathbf{k}}_{mathbf{*}}^{{top:}}{(mathbf{K}+{sigma:}_{n}^{2}{mathbf{I}}_{text{n}})}^{-1}{mathbf{k}}_{*},$$
    (3)
    where (:{mathbf{k}}_{*} = [k({text{x}}_{*} ,{text{x}}^{{left( 1 right)}} ), ldots :,k({text{x}}_{*} ,{text{x}}^{{left( N right)}} )]^{{{ top }:}}) is an N-dimensional vector, (:{k}_{**}=kleft({text{x}}_{*}{,text{x}}_{*}right)), and (:{mathbf{I}}_{text{n}}) is the N×N identity matrix. Note that the GPR not only offers pointwise estimations, (:{mu:}_{GP*})but also confidence estimates (:{sigma:}_{GPR*}^{2}). The relation between the input and the output is established as:$$:widehat{y}=fleft(mathbf{x}right)=sum:_{i=1}^{N}{alpha:}_{i}{k}_{theta:}left({mathbf{x}}^{left(iright)},mathbf{x}right)+{alpha:}_{0},$$
    (4)
    where αi is the weight assigned to each training predictor, α0 is the bias in the regression function, and kθ is the aforementioned covariance (kernel) function that evaluates the similarity between test and training data. In this paper, we selected the automatic relevance determination kernel:$$:Kleft({text{x}}_{i}{,text{x}}_{j}right)=nu::text{e}text{x}text{p}left(-sum:_{p=1}^{P}frac{{left({text{x}}_{i}^{left(pright)}-{text{x}}_{j}^{left(pright)}right)}^{2}}{{2sigma:}_{p}^{2}}right)+{sigma:}_{n}^{2}{delta:}_{ij},$$
    (5)
    where ν is a scaling factor, (:{sigma:}_{n}) accounts for the noise standard deviation, P is the number of predictors, and (:{sigma:}_{p}) is a dedicated parameter controlling the spread of the relations for each particular predictor p. The inverse of (:{sigma:}_{p}) represents the relevance of each predictor p. Model hyperparameters are collectively grouped in (:{varvec{theta:}=[nu:,sigma:}_{n,:}{sigma:}_{1},:dots:,:{sigma:}_{p}]) and model weights can be automatically optimized by maximizing the marginal likelihood in the training set33:$$:text{log:}pleft(mathbf{y}left|{mathbf{x}}_{text{i}},theta:right.right)=-frac{1}{2}{mathbf{y}}^{{top:}}{left(mathbf{K}+{sigma:}_{n}^{2}{mathbf{I}}_{text{n}}right)}^{-1}mathbf{y}-frac{1}{2}text{log:}left|mathbf{K}+{sigma:}_{n}^{2}{mathbf{I}}_{text{n}}right|-frac{N}{2}text{:log:}left(2pi:right).$$
    (6)
    Algorithm performance and CUE estimatesThe GPR model is evaluated with 2912 data pairs (8 predictors–CUE) obtained matching the in situ CUE data with the RS products. As the predictors exhibit a substantial range of variation among them, the inputs are scaled to 0–1 range to mitigate the risk of suboptimal performance. The calibration of the model is carried out by means of varying the number of training–validation data. Eleven cases of training–validation percentages are considered: 1% – 99%, 5% – 95%, 10% – 90%, 20% – 80%, 30% – 70%, 40% – 60%, 50% – 50%, 60% – 40%, 70% – 30%, 80% – 20%, and 90% – 10%. For every case, a hundred random initializations are conducted, and both the root mean square error (RMSE) and the coefficient of determination (R2) are computed to assess the GPR performance. Validation statistics were averaged across iterations, ensuring that all sites and years were represented in both training and testing datasets, thereby accounting for spatial and temporal variability. Subsequently, in view of the obtained results (see Results section), model evaluation was conducted using the 70%–30% split, enabling both biome-specific validation and assessment of predictor relevance according to the (:{sigma:}_{p}) values provided by the GPR. Finally, the GPR model (trained with all samples) is executed to obtain 1–km multitemporal CUE estimates at global scale from 2001 to 2023. This allows to report average CUE values and trends over different climate classes and biomes.Results and discussionModel performance over tower dataFigure 2 shows the assessment of the GPR model. The effect of model training/testing size reveals that for small training datasets, the GPR model does not perform optimally (Fig. 2(a)). Increasing the ratio of training/testing samples results in higher R² values and lower RMSE. The values are quite constant when datasets containing ≥ 70% of training samples are used for model training. Therefore, the 70% – 30% dataset is used to assess the accuracy of the CUE predictions provided by the GPR model. Figure 2(b) shows the assessment of the CUE predictions over the validation set composed of in situ EC data never used during model training. The accuracy metrics demonstrate strong agreement between the estimates and in situ measurements. Notably, the R² reaches 0.84 and the RMSE remains low (0.10), indicating strong predictive performance. The GPR CUE predictions reveal low mean error (ME = 0.01) and mean absolute error (MAE = 0.06). The GPR model provides the relevance of every predictor as the inverse of (:{sigma:}_{p}). Figure 2(c) shows this relevance normalized with the value of the maximum relevance predictor. The LAI is identified as the most relevant predictor for CUE, followed by the ET and LSTD. This is in accordance with other studies such as that by Liu et al.3. Through a hierarchical partitioning analysis between the explanatory variables and CUE, these authors found that LAI exerted greater impact than other factors related to climate (as temperature and precipitation). The accuracy metrics are disaggregated per biome type as shown in Table 2. The best results in all metrics are obtained over wetlands and evergreen needleleaf forests but must be biased by the low number of samples compared with the rest of biomes. The poorest agreement between estimates and in situ data is obtained over savannas.Fig. 2(a) GPR performance (RMSE and R2) as a function of the percentage of the training samples averaged over 100 random realizations for every training-validation data splitting. Shaded space indicates the standard deviation of the 100 realizations in every case. (b) CUE predicted by the GPR model (70% of the samples for training) over the in situ validation set (30% of the samples). (c) GPR input relevance for CUE. Relevance normalized with respect to the most relevant predictor.Full size imageTable 2 Performance of the GPR over the EC validation set per biome type. CODES according to IGBP land cover classification system34: CRO (crop), DBF (deciduous broadleaf forest), ENF (evergreen needleleaf forests), GRA (grassland), MF (mixed forest), OSH (open shrubland), SAV (savanna), WET (permanent wetland), WSA (woody savanna).Full size tableGlobal CUEThe execution of the GPR model provided estimates at a 1-km spatial resolution and 8-day temporal frequency, which were subsequently aggregated into annual means for the 2001–2023 period. In this section, the mean global CUE over this period is presented.The global mean CUE computed by the GPR model (0.43 ± 0.08) is, as expected, lower than the CUE’ obtained from MODIS (0.49 ± 0.06) during the same period. Similar underestimation behavior is found when compared with other studies that compute CUE’. For instance, Zhang et al.35 report a CUE’ value of 0.52 from MODIS data in the 2000–2003 period. He et al.36 employed five process-based models to estimate a global average CUE’ of 0.45 ± 0.05, slightly higher than the global mean obtained in the present study using GPR. In addition, the same study also reported a CUE’ value of 0.48 ± 0.05 from MODIS data in the 2000–2012 period, which is very similar to the one obtained for MODIS in the present study for the 2001–2023 period. Tang et al.37 and Jin et al.12 found a CUE’ value of 0.488 ± 0.136 and 0.50 ± 0.13, respectively.Figure 3 exhibits the global spatial distribution of the mean CUE, associated GPR pixel-wise uncertainty, and the spatial difference concerning CUE’ computed from the MOD17A3HGF product. A considerable global CUE variation, roughly ranging from 0.3 to 0.9 is found (Fig. 3, top). The sigma value provided by the GPR (Fig. 3, middle) ranges from 0.01 to 0.05. The spatial variation of CUE reported in Fig. 3 (top) highlights a clear latitudinal path, suggesting a dependence on climate variability. CUE increases with latitude: from minima around tropical zones (including tropical rainforests and tropical savannas), as also found by Jin et al.12 and Gang et al.38, to reaching maximum values in high northern latitudes (e. g., subarctic climate areas in Canada and Siberia). Street et al.39 analyzed the European subarctic region and showed the influence of mosses –that have CUE values greater than those observed for vascular plants– to increase the ecosystem CUE. The highest northern latitudes show CUE lower than the subarctic areas. Consistent with other global upscaling approaches, the sparse spatial distribution of flux tower data is a known source of uncertainty in global upscaling. The GPR framework directly addresses this limitation by providing explicit uncertainty, thereby identifying regions where model estimates are likely to be less reliable. The spatial pattern of the uncertainty is quite constant along the latitudes as shown in Fig. 3 (middle), with higher uncertainties over very high latitudes and some zones over the tropics.Figure 3 (bottom) shows the spatial difference ((:{varDelta:}_{text{C}text{U}text{E}})) between mean carbon use efficiencies obtained by the GPR model and MODIS. For the sake of clarity, from now on, CUEGPR and CUE’MODIS will refer to CUE and CUE’ values obtained by the GPR model and MODIS product (NPP and GPP from the MOD17A3HGF product), respectively. In general, a systematic CUEGPR < CUE’MODIS behavior is found, as expected by the different CUE definitions. However, there exist northern zones where very similar values are found. This can be partly due to very low soil heterotrophic respiration reported by other studies40,41 over northern zones, which reduces the differences between CUEs. In addition, our findings suggest CUEGPR ≥ CUE’MODIS and high uncertainty over these northern zones (see Fig. 3 (middle)). Other studies42 also reported largest uncertainty over high latitudes.The upscaling process is built upon the FLUXNET network, which has denser coverage in North America and Europe. To address this well-known limitation, the GPR model operates in the environmental feature space rather than geographic space. This approach leverages globally predictor variables to apply learned eco-physiological relationships by the GPR model. Despite the robustness of this methodology, the uncertainties provided by the GPR model could be larger in regions and biomes that are sparsely sampled. Expanding the in-situ monitoring network in these areas remains a critical priority for the global carbon cycle community. In addition, it is crucial to note that the CUEGPR framework is not intended as a numerical ‘correction’ for previous CUE’MODIS or similar estimates, but rather as a complementary metric that answers a different and broader ecological question. While CUE’ accurately reflects efficiency at the producer level, CUE evaluates the net outcome of carbon assimilation and total ecosystem respiration. The global patterns of CUE presented here therefore offer a new benchmark for the scientific community, particularly for evaluating and constraining the performance of Earth System Models, which must accurately simulate the complete ecosystem carbon balance. Although the GPR framework does not explicitly propagate uncertainties from flux tower observations and satellite inputs to the global scale —an interesting aspect that remains for future research— the dataset presented here contributes to improving our understanding of the terrestrial carbon sink. Both CUEGPR and CUE’MODIS are further analyzed in terms of the climatic areas according to the Köppen–Geiger climate classification43, and as a function of the biome type.Fig. 3Mean annual CUEGPR (top), uncertainty (middle), and mean annual differences between CUEGPR and CUE’MODIS (bottom) in the 2001–2023 period. The maps were generated with the Arcmap v.10.5 software (https://desktop.arcgis.com/es/arcmap/).Full size imageMean CUE for the different climatic classesGiven the latitudinal gradient observed in CUE values, the first analysis is in terms of the Köppen-Geiger climate classification, which was updated by Beck et al.43 using high-resolution, observation-based climatologies. Supplementary Table S1 available online summarizes the findings (mean annual values and standard deviation) over the five major climatic classes: tropical, arid (excluding main deserts, masked out in Fig. 3), temperate, cold and polar. As shown CUEGPR < CUE’MODIS for all the climatic classes, except for cold areas, where it shows a slightly higher value. In addition, Fig. 4 shows the distributions (represented using boxplots) of annual CUEGPR and CUE’MODIS. The edges of each box represent the first quartile (25th percentile) and the third quartile (75th percentile), while the whiskers extend to the 5th and 95th percentiles. The black line within the boxes stands for the median CUE value. Note that the median CUEGPR is always lower than the median CUE’MODIS, even in cold areas. Complementarily, results for the sub-classes according to the Köppen-Geiger climate classification are provided on Supplementary Figs. S3–S7 online. These kinds of comparisons between different carbon use efficiency definitions are carried out to illustrate the ecological significance of incorporating Rh. As dictated by their theoretical relationship (CUE = CUE’ – Rh/GPP), CUE must be less than or equal to CUE’. The reported results, therefore, do not validate the GPR model’s predictive power, but rather provide a quantitative assessment of the Rh/GPP ratio’s impact on carbon retention efficiency across different climate zones. The differences reveal how ecosystems belonging to different climate types vary in their capacity to retain assimilated carbon after accounting for all respiratory losses.Fig. 4CUE from the GPR model and CUE’ from MODIS during the 2001–2023 period over the major climate classes according to the Köppen-Geiger climate classification.Full size imageAs mentioned above, CUE values increase from tropical to cold climates and decrease again towards the polar zones. Concerning climatic differences during the 2001–2023 period, Supplementary Table S2 available online reports the mean CUEGPR and CUE’MODIS for every of the 30 sub-climatic classes. The values obtained by the GPR model are systematically lower than those reported by MODIS. However, over the temperate – dry winter and hot summer (Cwa), and cold – dry winter and hot summer (Dwa) climatic zones the mean values of both approaches are coincident (0.50 and 0.49, respectively). The highest mean CUEGPR values are found on temperate zones dominated by dry winter and cold summer (Cwc) (0.57 ± 0.06), and subarctic zones such as cold – dry and cold summer (Dsc) (0.57 ± 0.11), cold – dry summer and very cold winter (Dsd) (0.57 ± 0.08), and cold – dry and very cold winter (Dwd) (0.57 ± 0.04). In contrast, the lowest CUEGPR values are found over the tropical – rainforest climate class (Af) (0.35 ± 0.05). This similar behavior is shown by other studies12. Over the tundra polar climatic class, CUEGPR = 0.49 ± 0.03, which is very similar to 0.5 (computed as NEP/GPP) found by Reichle44.CUE mean values for the different biomesFor a given climatic zone, differences in CUE demonstrate that vegetation type has an obvious impact on CUE13. In fact, maintenance metabolism (Ra) costs are related to the biome type and to its adaptation to the environmental conditions. The biome type also determines the percentage of total ecosystem respiration that the Rh represents (lower for forests than for prairies and crops)6. Carbon fluxes, and therefore CUE values, are related to vegetation type. In this subsection, CUE values are analyzed as a function of the biome type (Fig. 5; Table 3). To account for the impact of land cover dynamics on CUE, we employed a year-specific methodology using the MODIS Land Cover Type (MCD12Q1) Version 6.1 product. For each year within our study period, we utilized the corresponding annual MCD12Q1 land cover map for that same year. For a given year, we calculated the CUE for every pixel. Then, using the land cover map for that year, we spatially aggregated these pixel-level CUE values based on the International Geosphere-Biosphere Programme (IGBP) classification scheme. This produced an annual time series of mean CUE for each IGBP biome. The overall aggregated results presented in this paper were then computed by combining these annual means over the entire 2001–2023 period. The analysis of temporal trends was conducted using this year-by-year time series of CUE for each biome. This dynamic approach ensures that our findings incorporate the influence of inter-annual land cover changes, such as those caused by deforestation, fires, and land use conversion, on ecosystem carbon dynamics. As mentioned in the introductory section, low CUE values imply that a little amount of carbon is converted to biomass and biological products, i.e., less carbon is retained in the organism and more is released4. It should be noted that, while climate drives the vegetation activity, a given vegetation type or biome can be found in areas showing different climatic conditions. Similar to the climatic zones results, median CUEGPR < median CUE’MODIS in all biomes (see Fig. 5). In OSH (open shrublands) and WET (wetlands), the whiskers of the boxplots suggest that, although the median CUEGPR is lower than CUE’MODIS, the mean value of the GPR could be equal or higher than that from MODIS, which is confirmed by values in Table 3.Fig. 5CUE from the GPR model and CUE’ from MODIS per biome type during the 2001–2023 period.Full size imageTable 3 Annual CUE values (mean and standard deviation) from the 2001–2023 period per biome type for the GPR and MODIS estimates. CODES according to IGBP land cover classification system34: CRO (crops), DBF (deciduous broadleaf forests), ENF (evergreen needleleaf forests), EBF (evergreen broadleaf forests), DNF (deciduous needleleaf forests), GRA (grasslands), MF (mixed forests), OSH (open shrublands), CSH (closed shrublands), SAV (savannas), WET (permanent wetlands), WSA (woody savannas).Full size tableIn relation to the CUE dependence on biome type, Jin et al.12 and Tang et al.37 found mean CUE’ values for CRO equal to 0.58 ± 0.12 and 0.566 ± 0.145, respectively. Luo et al.42 reported CUE’ values of 0.50 ± 0.09, 0.46 ± 0.10 and 0.39 ± 0.10 for CRO, DBF and ENF, respectively, and identical values for EBF and SAV (0.32 ± 0.12). Our results suggest higher values for ENF, EBF and SAV (CUEGPR = 0.47 ± 0.09, CUEGPR = 0.37 ± 0.04 and CUEGPR = 0.41 ± 0.08, respectively), but lower for CRO and DBF (CUEGPR = 0.42 ± 0.08 and CUEGPR = 0.43 ± 0.09). Tang et al.37 reported a CUE’ value of 0.464 ± 0.127 for forests, and Jin et al.12 0.44 ± 0.13 for MF, which are lower than the CUE reported in this study for MF (CUEGPR = 0.52 ± 0.09). Regarding GRA, CUEGPR = 0.44 ± 0.07, which is slightly lower than the CUE’ obtained by Tang et al.37 (0.457 ± 0.109).Our results show that, for CSH (closed shrublands), the mean CUEGPR = 0.42 ± 0.06 is clearly lower than that for OSH (open shrublands) (0.53 ± 0.12), indicating that the density or “openness” of the shrub canopy influences its CUE. For OSH, CUEGPR = 0.53 ± 0.12 is slightly higher than the mean CUE’ (:(sim)0.51) reported by Jin et al.12 and the one obtained in the present study with MODIS (CUE’MODIS = 0.48 ± 0.14). Table 3 shows that CUEGPR < CUE’MODIS, except for OSH, which is consistent with the spatial pattern found in CUE (previous section). In the case of savanna biomes, CUEGPR = 0.41 ± 0.08 and CUEGPR = 0.45 ± 0.11 for SAV (savannas) and for WSAV (woody savannas) are obtained, respectively. This suggests that savanna CUE may be lower due to the prevalence of C4 photosynthesis in grasslands, which implies higher energy and carbon costs compared to C3 photosynthesis45. Woody components in savannas tend to use the C3 pathway, which could contribute to higher CUE. Regarding WET (permanent wetlands), CUEGPR = 0.50 ± 0.10 is reported, which is lower than the mean CUE’ value found by Tang et al.37 (0.607 ± 0.133). However, our result is consistent with the CUE value of 0.504 over aquatic ecosystems reported by Reichle44.TrendThe computation of annual multitemporal estimates from 2001 to 2023 allows the provision of CUE trends at a global scale. All the trends (S) in this section have been quantified using the Theil-Sen slope estimator46,47,48. Note that all slopes are significant according to Hamed and Rao modified Mann-Kendall test49,50. SGPR and SMODIS will refer to the trend slopes obtained from CUEGPR and CUE’MODIS, respectively.Fig. 6Annual CUE trend (SGPR) for the 2001–2023 period. The map was generated with the Arcmap v.10.5 software (https://desktop.arcgis.com/es/arcmap/).Full size imageFigure 6 shows the pixel-wise trend (slope) of the annual CUEGPR in the 2001–2023 period at global scale. The mean slope is (:{S}_{text{GPR}}=left(-1.2pm:0.3right)times:{10}^{-2}:{text{d}text{e}text{c}text{a}text{d}text{e}}^{-1}). This reveals a decreasing global carbon use efficiency during the last two decades, which coincides with the decline rate values expected at the end of the 21 st century as estimated by Chen et al.16, mainly attributed to the temperature increase. Zhang et al.51 also revealed a negative trend for global CUE from 2000 to 2009. The spatial pattern of the trend obtained from 2001 to 2023 is in accordance with Gang et al.38, but disagrees from Zhang et al.51 in some areas (mainly South Africa). The discrepancy can be caused partly by the different and shorter temporal period used in that study (2000 to 2009), which highlights the importance of continuously assessing CUE to detect changes in its trend. The spatial pattern shows that the areas exhibiting significant positive changes ((:{S}_{text{GPR}}>0,:)in red color) are found in humid zones (as those in South America and in central-South Africa) as well as in arid zones (Australia and western USA) according to Zomer et al.52. In any case, an accelerated aridification has been identified by Sardans et al.53 in these zones, who enhanced central Africa as a new hotspot. Areas with negative trends ((:{S}_{text{GPR}}<0)) are also found in both humid and arid zones: the highest decreasing CUE is reported over the Amazon (humid), India, and northwestern Australia (more arid).Some of the above spatial patterns of negative trends (in South America, central Africa, and Southeast Asia) coincide with the forest carbon losses (including aboveground and belowground biomass carbon loss and soil organic carbon loss) identified by Feng et al.54 across the tropics. Most of them are attributed to agricultural practices and expansion (replacing the rainforests), confirming a dominant role of agriculture in long-term pan-tropical carbon reductions on formerly forested landscapes. Wang et al.55 show a net increase of the above-ground biomass carbon stock in Africa, but the rainforests present a loss. These authors highlight the human-induced deforestation and water stress (especially the vapor pressure deficit) as the most important variables explaining the spatial and temporal above-ground biomass variations. The negative CUE trend obtained in most part of Africa (not only in tropical forest but also in the transition zone to savanna woodlands and even in savannas) suggests the necessity of considering not only GPP but also respiration fluxes to further characterize the ecosystem state.In South America, a region with negative CUE trend is shown (Fig. 6), which mainly comprises tropical and subtropical forests (Amazonia), grasslands and savannas (for example, the north of the Brazilian Cerrado). The Brazilian Cerrado, which is considered the most biodiverse savanna on the planet, is affected by a degradation process due to both human activity (agricultural expansion and forest fires) and climate variability56. Simulated data by Delgado et al.57 show that tropical and subtropical forest, and savannas in South America will continue to show a decrease in vegetation activity due to the expected increase in air temperature as well as increasing fires during the dry season. Globally, both positive and negatives trends span different climatic zones and very different ecosystems, which is analyzed in the next sections.CUE trend for the different climatic classesAn analysis per climate type (Fig. 7; Table 4) revealed decreasing trends over all major climatic classes except in arid zones where low positive trend is found. The negative trends are reported for both the GPR model and MODIS ((:{S}_{text{GPR}}<0,:{:S}_{text{MODIS}}<0)), however, over temperate zones (:left|{S}_{text{GPR}}right|>left|{S}_{text{MODIS}}right|).Table 4 Trend slope of annual CUE and its corresponding p-value for the major Climatic classes (Fig. 7). SGPR and SMODIS refer to the trend slopes obtained from CUEGPR and CUE’MODIS, respectively.Full size tableComplementarily, online Supplementary Figs. S8 to S12 show the CUE trend reported for all 30 climatic classes. All three tropical sub-types follow a negative trend (see Supplementary Fig. S8 online), whereas in the case of the arid sub-types the arid – steppe cold (BSk) shows a negative trend contrary to the rest of arid zones (see Supplementary Fig. S9 online). In the case of temperate sub-types there is a mixture of behaviors in trends (see Supplementary Fig. S10 online) that makes the final trend of temperate zones to be slightly negative. The cold climatic zone is largely dominated by the sub-type cold – no dry season and cold summer (Dfc) (see Supplementary Fig. S1 online), which makes that the Dfc trend (see Supplementary Fig. S11 online) has a major impact on the overall cold trend. Finally, the polar CUE trend is mainly driven by the sub-type polar – tundra (ET) as shown in Supplementary Fig. S12 online.Fig. 7Trend of the mean annual CUE during the 2001–2023 period over the major climatic classes, as obtained by the GPR model and MODIS (CUE’). All trends are significant as reported by the p-values (pGPR and pMODIS).Full size imageCUE trend for the different biomesFigure 8 shows the temporal evolution of CUEGPR and CUE’MODIS per biome type. In general, the CUEGPR temporal evolution resulted in slightly noisier temporal estimates than CUE’MODIS. This can be also observed in the slope’s error, being generally higher in the case of the GPR model (see Table 5). These interannual variabilities are usually induced by machine learning models in long-term trends of carbon uptake58. Both CUEGPR and CUE’MODIS show positive trends ((:{S}_{text{GPR}}>0,:{:S}_{text{MODIS}}>0)) for SAV, WSAV, and CRO, and negative trends ((:{S}_{text{GPR}}<0,:{:S}_{text{MODIS}}<0)) for DBF, EBF, ENF, DNF, GRASS, OSH, and CSH. In the case of MF (mixed forests), (:{S}_{text{GPR}}<0), while MODIS shows a very slight positive trend ((:{:S}_{text{MODIS}}>0)) but two order of magnitude lower in absolute value ((:sim)10–6 yr–1 vs. 10–4 yr–1). For WET, the trends reported by the two models are opposite, but also low: (:{S}_{text{GPR}}sim)10–5 yr–1, and (:{:S}_{text{MODIS}}sim) − 10–6 yr–1). Both (:{S}_{text{GPR}})and (::{S}_{text{MODIS}}) show the highest negative values over evergreen forests (ENF and EBF). The highest positive trend in CUEGPR is found over croplands (CRO), whereas the highest positive trend in CUEMODIS is found over wooded savannas. There is a lack of studies reporting CUE trends over different biomes at global scale, and only few studies address this subject. For instance, Yang et al.59 find positive CUE trends over SAV and WSAV, and a negative trend for non-woody grasslands at global scale during the period 2000–2013. Du et al.60 report a negative CUE trend of (:sim) − 5 × 10−4 yr−1 for GRASS in the 2001–2017 period over the Ningxia province (northwest China), which is similar to the trend reported in this study at global scale (see Fig. 8). Lei et al.61 evaluate the 2001–2023 period and show a negative CUE’ trend for forests, grasslands, and croplands, and a positive trend over shrublands in the Nanling Mountains (700 km × 400 km region in China).Fig. 8Trend, as characterized by S, of the mean annual CUE per biome type during the 2001–2023 period. All trends are significant (p < 0.01) as reported by the pGPR and pMODIS.Full size imageTable 5 Trend slope of annual CUE and its corresponding p-value for the major Climatic classes (Fig. 7). SGPR and SMODIS refer to the trend slopes obtained from CUEGPR and CUE’MODIS, respectively.Full size tableConclusionsHaving efficient tools to characterize and quantify the functioning of ecosystems at a global scale is critical for climate change adaptation. The main indicator used by the plant community is based on autotrophic CUE’, a metric of plant-level efficiency that does not account for subsequent carbon losses due to heterotrophic respiration. This omission can lead to an overestimation of the efficiency with which entire ecosystems retain carbon. The present study bridges this gap by calculating and presenting the first global dataset of an ecosystem CUE, providing a robust measure of global terrestrial carbon sequestration efficiency. This study develops a potent, robust tool to map ecosystem carbon use efficiency at global scale, which allows to quantify CUE changes over time. The approach relies on an upscaling data-driven methodology combining in situ data with remote sensing observations into a GPR regression algorithm to retrieve global CUE. Results provide strong evidence of the algorithm’s high performance assessed over in situ data never used during model training.The GPR model fed with 1-km multitemporal annual data from 2001 to 2023 allows to provide a global mean annual CUE map, whose overall spatial pattern is in accordance with the studies related to atmosphere-biosphere carbon exchanges reported by the scientific community. Results show that the mean CUE values increase from tropical to cold climates and decrease again towards the polar zones. The findings revealed that the mean spatial-temporal CUE estimates obtained by the GPR model are generally lower than the mean CUE’ of MODIS reported by the present paper. This can be mainly explained by different definitions of carbon use efficiency: CUE from GPR model is defined as NEP/GPP, whereas MODIS CUE’ accounts for NPP/GPP. This distinction is crucial when comparing values from different approaches that use varying definitions for carbon use efficiency. A preliminary analysis of estimated mean CUE per biome type supports the hypothesis that the proportion and type of woody vegetation in savannas significantly influences its CUE, with woody savannas potentially being more efficient due to different photosynthetic pathways or carbon allocation strategies compared to non-woody savannas. This highlights the importance of accurately classifying and distinguishing between subtypes of savanna, as well as different shrubs, in remote sensing studies to understand biome-specific carbon dynamics.The GPR model reported a negative trend in the global mean CUE value, with a slope (decline rate) equal to (:left(-1.2pm:0.3right)times:{10}^{-3}:{text{y}text{r}}^{-1}), which is in accordance with future scenarios of global carbon use efficiency. The reported negative global CUE trend suggests a decreasing ecosystems’ capacity to sequester atmospheric CO2, which highlights the importance of assessing the terrestrial carbon uptake in the future.

    Data availability

    The datasets used and analysed during the current study are available from the corresponding author on reasonable request.
    ReferencesMigliavacca, M. et al. The three major axes of terrestrial ecosystem function. Nature 598, 468–472. https://doi.org/10.1038/s41586-021-03939-9 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bradford, M. A. & Crowther, T. W. Carbon use efficiency and storage in terrestrial ecosystems (Commentary). New. Phytol. 199, 7–9. https://doi.org/10.1111/nph.12334 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Liu, Z. & Han, L. Ecosystem carbon use efficiency in ecologically vulnerable areas in china: variation and influencing factors. Front. Plant. Sci. 13, 1062055. https://doi.org/10.3389/fpls.2022.1062055 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Manzoni, S. et al. Reviews and syntheses: carbon use efficiency from organisms to ecosystems—definitions, theories, and empirical evidence. Biogeosciences 15, 5929–5949. https://doi.org/10.5194/bg-15-5929-2018 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Sinsabaugh, R. L. et al. Plant, microbial and ecosystem carbon use efficiencies interact to stabilize microbial growth as a fraction of gross primary production. New. Phytol. 214, 1518–1526. https://doi.org/10.1111/nph.14485 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Reichle, D. E. Energy flow in ecosystems, in The global carbon cycle and climate change. Chap 8, 119–156. https://doi.org/10.1016/B978-0-12-820244-9.00008-1 (2020).Article 

    Google Scholar 
    Campos-Taberner, M. et al. Global carbon fluxes using multioutput Gaussian processes regression and MODIS products. IEEE J. Sel. Top. Appl. Earth Obs. 17, 11310–11321. https://doi.org/10.1109/JSTARS.2024.3413184 (2024).Article 

    Google Scholar 
    Nelson, J. A. et al. X-BASE: the first terrestrial carbon and water flux products from an extended data-driven scaling framework, FLUXCOM-X. Biogeosciences 21, 5079–5115. https://doi.org/10.5194/bg-21-5079-2024 (2024).Article 
    ADS 

    Google Scholar 
    Baldocchi, D. et al. Inter-annual variability of net and gross ecosystem carbon fluxes: A review. Agr For. Meteorol. 249, 520–533. https://doi.org/10.1016/j.agrformet.2017.05.015 (2018).Article 

    Google Scholar 
    Pandey, V., Harde, S., Rajasekaran, E. & Kumar Deb Burman, P. Gross primary productivity of terrestrial ecosystems: a review of observations, remote sensing, and modelling studies over South Asia. Theor. Appl. Climatol. 155, 8461–8491. https://doi.org/10.1007/s00704-024-05158-4 (2024).Article 
    ADS 

    Google Scholar 
    Fernández-Martínez, M. et al. Spatial variability and controls over biomass stocks, carbon fluxes and resource-use efficiencies in forest ecosystems. Trees Struct. Funct. 28, 597–611. https://doi.org/10.1007/s00468-013-0975-9 (2014).Article 
    CAS 

    Google Scholar 
    Jin, C. et al. Spatiotemporal variation in carbon use efficiency derived from eddy-covariance measurement of global terrestrial biomes. Agr For. Meteorol. https://doi.org/10.1016/j.agrformet.2024.110318 (2025).Article 

    Google Scholar 
    Chen, Z. & Yu, G. Spatial variations and controls of carbon use efficiency in China’s terrestrial ecosystems. Sci. Rep. https://doi.org/10.1038/s41598-019-56115-5 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    DeLucia, E. H. et al. Forest carbon use efficiency: is respiration a constant fraction of gross primary production? Global Change Biol. 13, 1157–1167. https://doi.org/10.1111/j.1365-2486.2007.01365.x (2007).Article 
    ADS 

    Google Scholar 
    Qian, L. et al. Assessing and improving the high uncertainty of global gross primary productivity products based on deep learning under extreme climatic conditions. Sci. Total Environ. https://doi.org/10.1016/j.scitotenv.2024.177344 (2024).Article 
    PubMed 

    Google Scholar 
    Chen, G. et al. Changes in global carbon use efficiency in the 21st century and the potential controlling factors. Environ. Res. Lett. https://doi.org/10.1088/1748-9326/ad3e90 (2024).Article 

    Google Scholar 
    Letchov, G. Carbon-use efficiency of terrestrial ecosystems under stress condition in South East Europe (MODIS; NASA). Proceedings 2 (363), 103390ecrs–103390ecr2 (2018).
    Google Scholar 
    Desai, A. R. et al. Drivers of decadal carbon fluxes across temperate ecosystems. J. Geophys. Res. Biogeosci. 127, e2022JG007014. https://doi.org/10.1029/2022JG007014 (2022).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Baldocchi, D. et al. FLUXNET: a new tool to study the Temporal and Spatial variability of ecosystem-scale carbon dioxide, water vapor, and energy flux densities. Bull. Am. Meteorol. Soc. 82, 2415–2434 (2001).2.3.CO;2″ data-track-item_id=”10.1175/1520-0477(2001)082<2415:FANTTS>2.3.CO;2″ data-track-value=”article reference” data-track-action=”article reference” href=”https://doi.org/10.1175%2F1520-0477%282001%29082%3C2415%3AFANTTS%3E2.3.CO%3B2″ aria-label=”Article reference 19″ data-doi=”10.1175/1520-0477(2001)082<2415:FANTTS>2.3.CO;2″>Article 
    ADS 

    Google Scholar 
    Zeng, J. et al. Global terrestrial carbon fluxes of 1999–2019 estimated by upscaling eddy covariance data with a random forest. Sci. Data https://doi.org/10.1038/s41597-020-00653-5 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Papale, D. & Valentini, R. A new assessment of European forests carbon exchanges by eddy fluxes and artificial neural network spatialization. Glob Change Biol. 9, 525–535. https://doi.org/10.1046/j.1365-2486.2003.00609.x (2003).Article 
    ADS 

    Google Scholar 
    Reyes-Muñoz, P. et al. Tower-to-global upscaling of terrestrial carbon fluxes driven by MODIS-LAI, Sentinel-3-LAI and ERA5-Land data. Ecol. Indic. 177 (11359), 7. https://doi.org/10.1016/j.ecolind.2025.113597 (2025).Article 
    CAS 

    Google Scholar 
    Pastorello, G. et al. The FLUXNET2015 dataset and the oneflux processing pipeline for eddy covariance data. Sci. Data. 7, 225. https://doi.org/10.1038/s41597-020-0534-3 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Reichstein, M. et al. On the separation of net ecosystem exchange into assimilation and ecosystem respiration: review and improved algorithm. Glob Change Biol. 11, 1424–1439. https://doi.org/10.1111/j.1365-2486.2005.001002.x (2005).Article 
    ADS 

    Google Scholar 
    García-Haro, F. J. et al. Derivation of global vegetation biophysical parameters from EUMETSAT Polar system. ISPRS J. Photogramm Remote Sens. 139, 57–74. https://doi.org/10.1016/j.isprsjprs.2018.03.005 (2018).Article 
    ADS 

    Google Scholar 
    Liang, S. & Wang, D. Moderate resolution imaging spectroradiometer (MODIS) downward shortwave radiation (MCD18A1) and photosynthetically active radiation (MCD18A2. Algorithm Theoretical Basis Doc. https://doi.org/10.5067/MODIS/MCD18C2.061 (2017).Article 

    Google Scholar 
    Hulley, G. C., Hook, S. J. & Hughes, C. MODIS MOD21 land surface temperature and emissivity algorithm theoretical basis document. Pasadena, CA: Jet Propulsion Laboratory, National Aeronautics and Space Administration. (2012). https://emissivity.jpl.nasa.gov/downloads/examples/documents/MOD21_LSTE_ATBD_Hulley_v2.0_20121116_pmbv1.pdfMu, Q., Zhao, M. & Running, S. W. MODIS global terrestrial evapotranspiration (ET) product (NASA MOD16A2/A3). Algorithm Theoretical Basis Doc. Collect. 5, 600. https://doi.org/10.5067/MODIS/MOD16A2.006 (2013).Article 

    Google Scholar 
    Knyazikhin, Y. et al. MODIS leaf area index (LAI) and fraction of photosynthetically active radiation absorbed by vegetation (FPAR) product (MOD15) algorithm theoretical basis document. (1999). https://doi.org/10.5067/MODIS/MCD15A3H.061Camps-Valls, G. et al. A unified vegetation index for quantifying the terrestrial biosphere. Sci. Adv. 7, eabc7447. https://doi.org/10.1126/sciadv.abc7447 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Strahler, A. H. et al. MODIS BRDF/albedo product: algorithm theoretical basis document version 5.0, 1–53 (1999). https://doi.org/10.5067/MODIS/MCD43A4.061Running, S. & Zhao, M. MOD17A3HGF MODIS/Terra net primary production Gap-Filled yearly L4 global 500 m SIN grid V006. (2019). https://doi.org/10.5067/MODIS/MOD17A3HGF.006Rasmussen, C. E. & Williams, C. K. I. Gaussian Processes for Machine Learning (The MIT Press, 2006).Loveland, T. R. et al. Development of a global land cover characteristics database and IGBP discover from 1 Km AVHRR data. Int. J. Remote Sens. 21, 1303–1330. https://doi.org/10.1080/014311600210191 (2000).Article 

    Google Scholar 
    Zhang, Y. J. et al. Global pattern of NPP to GPP ratio derived from MODIS data: effects of ecosystem type, geographical location and climate. Global Ecol. Biogeogr. 18, 280–290. https://doi.org/10.1111/j.1466-8238.2008.00442.x (2009).Article 

    Google Scholar 
    He, Y. et al. Global patterns of vegetation carbon use efficiency and their climate drivers deduced from MODIS satellite data and process-based models. Agric. For. Meteorol. 256–257, 150–158. https://doi.org/10.1016/j.agrformet.2018.03.009 (2018).Article 
    ADS 

    Google Scholar 
    Tang, et al. Global variability of carbon use efficiency in terrestrial ecosystems. Biogeosci. Discuss, 1–19. https://doi.org/10.5194/bg-2019-37 (2019).Gang, C. et al. Divergent responses of terrestrial carbon use efficiency to climate variation from 2000 to 2018. Glob Planet. Change. 208, 103709. https://doi.org/10.1016/j.gloplacha.2021.103709 (2022).Article 

    Google Scholar 
    Street, L. E. et al. The role of mosses in carbon uptake and partitioning in artic vegetation. New. Phytol. 199, 163–175. https://doi.org/10.1111/nph.12285 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kurganova, I. N. et al. A pilot National network for monitoring soil respiration in russia: first results and prospects of development. Dokl. Earth Sc. 519, 1947–1954. https://doi.org/10.1134/S1028334X24603377 (2024).Article 
    ADS 
    CAS 

    Google Scholar 
    Mukhortova, L. et al. Respiration of Russian soils: Climatic drivers and response to climate change. Sci. Total Environ. 785, 147314. https://doi.org/10.1016/j.scitotenv.2021.147314 (2021).Article 
    CAS 

    Google Scholar 
    Luo, X. et al. Global variation in vegetation carbon use efficiency inferred from eddy covariance observations. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-025-02753-0 (2025).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Beck, H. E. et al. High-resolution (1 km) Köppen–Geiger maps for 1901–2099 based on constrained CMIP6 projections. Sci. Data. 10, 724. https://doi.org/10.1038/s41597-023-02549-6 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Reichle, D. E. The Global Carbon Cycle and Climate Change: Scaling Ecological Energetics from Organism To the Biosphere (Elsevier, 2023).Naidu, K. R. et al. Dark respiration of leaves in selected C4 and C3 tropical weed species. Z. Pflanzenphysiol. 99, 85–88. https://doi.org/10.1016/S0044-328X(80)80116-4 (1980).Article 
    CAS 

    Google Scholar 
    Sen, P. K. Estimates of the regression coefficient based on kendall’s Tau. J. Am. Stat. Assoc. 63, 1379–1389. https://doi.org/10.1080/01621459.1968.10480934 (1968).Article 
    MathSciNet 

    Google Scholar 
    Theil, H. A rank-invariant method of linear and polynomial regression analysis. Indag Math. 12, 386–392. https://doi.org/10.1007/978-94-011-2546-8_20 (1950).Article 
    MathSciNet 

    Google Scholar 
    Fernandes, R. & Leblanc, S. G. Parametric (modified least squares) and non-parametric (Theil–Sen) linear regressions for predicting biophysical parameters in the presence of measurement errors. Remote Sens. Environ. 95, 303–316. https://doi.org/10.1016/j.rse.2005.01.005 (2005).Article 
    ADS 

    Google Scholar 
    Mann, H. B. Nonparametric tests against trend. Econometrica 13, 245–259. https://doi.org/10.2307/1907187 (1945).Article 
    MathSciNet 

    Google Scholar 
    Kendall, M. G. Rank Correlation Methods (Charles Grifin, 1975).Zhang, Y. et al. Climate-driven global changes in carbon use efficiency. Glob Ecol. Biogeogr. 23 (2), 144–155. https://doi.org/10.1111/geb.12086 (2014).Article 

    Google Scholar 
    Zomer, R. J., Xu, J. & Trabucco, A. Version 3 of the global aridity index and potential evapotranspiration. Database Sci. Data. 9, 409. https://doi.org/10.1038/s41597-022-01493-1 (2022).Article 
    PubMed 

    Google Scholar 
    Sardans, J. et al. Growing aridity poses threats to global land surface. Commun. Earth Environ. 5, 776. https://doi.org/10.1038/s43247-024-01935-1 (2024).Article 
    ADS 

    Google Scholar 
    Feng, Y. et al. Doubling of annual forest carbon loss over the tropics during the early twenty-first century. Nat. Sustain. 5, 444–451. https://doi.org/10.1038/s41893-022-00854-3 (2022).Article 

    Google Scholar 
    Wang, M. et al. Satellite observed aboveground carbon dynamics in Africa during 2003–2021. Remote Sens. Environ. 301, 113927. https://doi.org/10.1016/j.rse.2023.113927 (2024).Article 

    Google Scholar 
    dos Lopes, G. et al. Degradation of the Brazilian cerrado: interactions with human disturbance and environmental variables. Ecol. Manage. 482, 118875. https://doi.org/10.1016/j.foreco.2020.118875 (2021).Article 

    Google Scholar 
    Delgado, R. C. et al. Degradation of South American biomes: what to expect for the future? Environ. Impact Assess. Rev. 96, 106815. https://doi.org/10.1016/j.eiar.2022.106815 (2022).Article 

    Google Scholar 
    Jung, M. et al. Scaling carbon fluxes from eddy covariance sites to globe: synthesis and evaluation of the FLUXCOM approach. Biogeosciences 17, 1343–1365. https://doi.org/10.5194/bg-17-1343-2020 (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Yang, Y. et al. Assessing the spatio-temporal dynamic of global grassland carbon use efficiency in response to climate change from 2000 to 2013. Acta Oecol. 81, 22–31. https://doi.org/10.1016/j.actao.2017.04.004 (2017).Article 
    ADS 

    Google Scholar 
    Du, L. et al. Carbon use efficiency of terrestrial ecosystems in desert/grassland biome transition zone: A case in Ningxia province, Northwest China. Ecol. Indic. 120, 106971. https://doi.org/10.1016/j.ecolind.2020.106971 (2021).Article 
    CAS 

    Google Scholar 
    Lei, S. et al. Spatiotemporal variation in carbon and water use efficiency and their influencing variables based on remote sensing data in the Nanling mountains region. Remote Sens. 17 (4), 648. https://doi.org/10.3390/rs17040648 (2025).Article 
    ADS 

    Google Scholar 
    Download referencesAcknowledgementsThis study was supported by Grant PID2020-118036RB-I00 funded by MCIN/AEI/10.13039/501100011033, and by LSA-SAF (EUMETSAT).FundingThis study was funded by Spanish Ministry of Science, Innovation and Universities MCIN/AEI/10.13039/501100011033 with Grant PID2020-118036RB-I00.Author informationAuthors and AffiliationsEnvironmental Remote Sensing group (UV-ERS), Universitat de València, Dr. Moliner, Burjassot, València, 46100, SpainM. Campos-Taberner, M. A. Gilabert, S. Sánchez-Ruiz, B. Martínez, A. Jiménez-Guisado & F. J. García-HaroAuthorsM. Campos-TabernerView author publicationsSearch author on:PubMed Google ScholarM. A. GilabertView author publicationsSearch author on:PubMed Google ScholarS. Sánchez-RuizView author publicationsSearch author on:PubMed Google ScholarB. MartínezView author publicationsSearch author on:PubMed Google ScholarA. Jiménez-GuisadoView author publicationsSearch author on:PubMed Google ScholarF. J. García-HaroView author publicationsSearch author on:PubMed Google ScholarContributionsM.C-T. and M.A.G. planned the research and wrote the manuscript. M.C-T., S.S-R. and B.M. gathered data. A.J-G. and F.J.G-H. contributed to data analysis and paper development.Corresponding authorCorrespondence to
    M. Campos-Taberner.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleCampos-Taberner, M., Gilabert, M.A., Sánchez-Ruiz, S. et al. Ecosystem carbon use efficiency at global scale from upscaling eddy-covariance data with machine learning and MODIS products.
    Sci Rep 15, 44496 (2025). https://doi.org/10.1038/s41598-025-28151-xDownload citationReceived: 25 July 2025Accepted: 07 November 2025Published: 24 December 2025Version of record: 24 December 2025DOI: https://doi.org/10.1038/s41598-025-28151-xShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsCarbon use efficiency (CUE)Machine learningEddy-covariance dataMODISSpatiotemporal variability More

  • in

    Assessment of stability and yield performance of onion (Allium cepa L.) genotypes across diverse Indian environments

    AbstractOnion productivity in India is strongly influenced by genotype × environment interaction (GEI), complicating the identification of stable and high-yielding cultivars. Nineteen onion genotypes were evaluated in multi-environment trials across six diverse locations during two consecutive kharif seasons (2023–24), with locations treated as six distinct environments (E1-E6). Combined ANOVA revealed significant effects of genotypes, environments, and their interactions, indicating substantial GEI. AMMI analysis identified RO-1771 (G15), RO-1768 (G13), and RO-1774 (G17) as the most stable and high-yielding genotypes, while Junagadh (E3) was the most representative and discriminating environment. The first two principal components of the GGE biplot explained 83.86% of total GEI variation, efficiently capturing interaction patterns. Among the genotypes, ‘Bhima Dark Red’ (G18) consistently showed superior yield and stability, followed by RO-1768 and RO-1774, which also demonstrated broad adaptability. Junagadh was identified as an ideal site for both cultivation and varietal testing. Collectively, the GGE and AMMI analyses revealed that Bhima Dark Red (G18), RO-1768 (G13), and RO-1774 (G17) combined high yield with stability across multiple environments. These findings provide valuable guidance for onion breeding programs and support the development of cultivars adapted to diverse Indian agro-climatic conditions.

    Similar content being viewed by others

    AMMI and GGE biplot analysis for yield performance and stability assessment of selected Bambara groundnut (Vigna subterranea L. Verdc.) genotypes under the multi-environmental trials (METs)

    Article
    Open access
    23 November 2021

    Evaluation of genotype by environment interaction and adaptability in lowland irrigated rice hybrids for grain yield under high temperature

    Article
    Open access
    04 August 2021

    Envirotype-based delineation of environmental effects and genotype × environment interactions in Indian soybean (Glycine max, L.)

    Article
    Open access
    21 May 2024

    Data availability

    The authors confirm that the data supporting the findings of this study are available within the manuscript and/or in its supplementary material.
    ReferencesEtana, M. B., Aga, M. C. & Fufa, B. O. Major onion (Allium cepa L.) production challenges in Ethiopia: A review. J. Biol. Agric. Healthc. 9, 42–47. https://doi.org/10.7176/JBAH (2019).
    Google Scholar 
    Biswas, P., Das, S., Bar, A., Maity, T. K. & Mandal, A. R. Effect of micronutrient application on vegetative growth and bulb yield attributes of rabi onion (Allium cepa L.). Int. J. Curr. Microbiol. Appl. Sci. 9, 556–565. https://doi.org/10.20546/ijcmas.2020.903.065 (2020).
    Google Scholar 
    Colina-Coca, C., Ancos, B. & Sanchez-Moreno, C. Nutritional composition of processed onion: S-alk(en)yl-L-cysteine sulfoxides, organic acids, sugars, minerals and vitamin C. Food Bioprocess Technol. 7, 289–298. https://doi.org/10.1007/s11947-013-1150-4 (2014).
    Google Scholar 
    Slimestad, R., Fossen, T. & Vagen, I. Onions: A source of unique dietary flavonoids. J. Agric. Food Chem. 55, 10067–10080. https://doi.org/10.1021/jf0712503 (2007).
    Google Scholar 
    FAOSTAT. Onion production, area and productivity. FAO Statistics (2023). Available at: http://faostat3.fao.org/browse/Q/QC/E (accessed 10 July 2023).Gupta, A. J., Kaldate, S., Volaguthala, S. & Mahajan, V. Onion nutritional and nutraceutical composition and therapeutic potential of its phytochemicals assessed through preclinical and clinical studies. J. Funct. Foods 129, 106889. https://doi.org/10.1016/j.jff.2025.106889 (2025).
    Google Scholar 
    Gupta, A. J. et al. Studies on biochemical and nutraceutical profiling of 43 Indian onion (Allium cepa L.) genotypes. J. Food Meas. Charact. https://doi.org/10.1007/s11694-025-03425-0 (2025).
    Google Scholar 
    Khar, A. & Singh, H. Rapid methods for onion breeding. In Accelerated Plant Breeding (eds Gosal, S. S. & Wani, S. H.) 77–99 (Springer Nature, 2020). https://doi.org/10.1007/978-3-030-47298-6_4.
    Google Scholar 
    Romagosa, I., Borras-Gelonch, G., Slafer, G. & van Eeuwijk, F. Genotype by environment interaction and adaptation. In Sustainable Food Production 846–870 (Springer, 2013). https://doi.org/10.1007/978-1-4614-5797-8_199Malosetti, M., Ribaut, J. M. & van Eeuwijk, F. A. The statistical analysis of multi-environment data: Modeling genotype-by-environment interaction and its genetic basis. Front. Physiol. 4, 44. https://doi.org/10.3389/fphys.2013.00044 (2013).
    Google Scholar 
    Romagosa, I. & Fox, P. N. Genotype × environment interaction and adaptation. In Plant Breeding: Principles and Prospects 373–390 (Springer, 1993).Annicchiarico, P. Genotype × environment interactions – challenges and opportunities for plant breeding and cultivar recommendations. FAO, Rome, Italy (2002). Available at: http://www.fao.org/docrep/005/y4391e/y4391e00Yan, W. K., Kang, M. S., Ma, B., Woods, S. & Cornelius, P. L. GGE biplot vs. AMMI analysis of genotype-by-environment data. Crop Sci. 47, 643–655. https://doi.org/10.2135/cropsci2006.06.0374 (2007).
    Google Scholar 
    Yan, W. K. GGE biplot – A windows application for graphical analysis of multi-environment trial data and other types of two-way data. Agron. J. 93, 1111–1118. https://doi.org/10.2134/agronj2001.9351111x (2001).
    Google Scholar 
    Yan, W., Hunt, L. A., Sheng, Q. & Szlavnics, Z. Cultivar evaluation and mega-environment investigation based on GGE biplot. Crop Sci. 40, 597–605. https://doi.org/10.2135/cropsci2000.403597x (2000).
    Google Scholar 
    Crossa, J. Statistical analysis of multilocation trials. Adv. Agron. 44, 55–85. https://doi.org/10.1016/S0065-2113(08)60818-4 (1990).
    Google Scholar 
    Gupta, A. J., Benke, A. P., Mahajan, V., Chauhan, H. & Singh, M. Assessment of genetic diversity and stability performance of 38 genotypes of onion (Allium cepa L.). J. Hortic. Sci. Biotechnol. 99, 560–569. https://doi.org/10.1080/14620316.2024.2315941 (2024).
    Google Scholar 
    Gupta, A. J. et al. Assessing onion genotypes stability and potential in diverse Indian environments. Cogent Food Agric. 10, 2360606. https://doi.org/10.1080/23311932.2024.2360606 (2024).
    Google Scholar 
    Tahir, M. et al. GGE biplot an effective tool to study genotype and genotype × environment interaction; a case study in onion (Allium cepa L.). Pak. J. Agric. Sci. 57, 1565–1571 (2020).
    Google Scholar 
    Gabriel, K. R. The biplot graphic display of matrices with application to principal component analysis. Biometrika 58, 453–467. https://doi.org/10.1093/biomet/58.3.453 (1971).
    Google Scholar 
    Zobel, R. W., Wright, M. J. & Gauch, H. G. Statistical analysis of a yield trial. Agron. J. 80, 388–393. https://doi.org/10.2134/agronj1988.00021962008000030002x (1988).
    Google Scholar 
    Gauch, H. G. Statistical Analysis of Regional Yield Trials: AMMI Analysis of Factorial Designs (Elsevier, 1992).
    Google Scholar 
    Cooper, M. & Hammer, G. L. (eds) Plant Adaptation and Crop Improvement. CABI Publishing, Wallingford, UK; ICRISAT, Patancheru, India; and IRRI, Manila, Philippines (1996).Yan, W. & Tinker, N. A. Biplot analysis of multi-environment trial data: Principles and applications. Can. J. Plant Sci. 86, 623–645. https://doi.org/10.4141/P05-169 (2006).
    Google Scholar 
    Kaya, Y., Akcura, M. & Taner, S. GGE-biplot analysis of multi-environment yield trials in bread wheat. Turk. J. Agric. For. 30, 325–337 (2006).
    Google Scholar 
    Hagos, H. G. & Abay, F. AMMI and GGE biplot analysis of bread wheat genotypes in the northern part of Ethiopia. J. Plant Breed. Genet. 1, 12–18 (2013).
    Google Scholar 
    Hossain, M. A. et al. Integrating BLUP, AMMI, and GGE models to explore GE interactions for adaptability and stability of winter lentils (Lens culinaris Medik.). Plants 12, 2079. https://doi.org/10.3390/plants12112079 (2023).
    Google Scholar 
    Omrani, A. et al. Evaluation of grain yield stability in some selected wheat genotypes using AMMI and GGE biplot methods. Agronomy 12, 1130. https://doi.org/10.3390/agronomy12051130 (2022).
    Google Scholar 
    Wang, R. et al. Assessment of yield performances for grain sorghum varieties by AMMI and GGE biplot analyses. Front. Plant Sci. 14, 1261323. https://doi.org/10.3389/fpls.2023.1261323 (2023).
    Google Scholar 
    Stansluos, A. A. L. et al. Genotype × trait (GT) biplot analysis for yield and quality stability in some sweet corn (Zea mays L. Saccharata Sturt.) genotypes. Agronomy 13, 1538. https://doi.org/10.3390/agronomy13061538 (2023).
    Google Scholar 
    Hnizil, O. et al. Dissecting genotype-by-environment interactions in Moroccan wheat: An advanced biplot and heatmap analysis unveiling agronomic, quality traits, and genotypic stability for tailored breeding strategies. Plants 13, 1068. https://doi.org/10.3390/plants13081068 (2024).
    Google Scholar 
    Sultana, F. et al. Heatmap clustering and performance analysis of cotton genotypes in response to environmental conditions. Sci. Rep. 15, 19297. https://doi.org/10.1038/s41598-025-02682-9 (2025).
    Google Scholar 
    Li, Z., Zheng, B. & He, Y. Understanding the effects of growing seasons, genotypes, and their interactions on the anthesis date of wheat sown in North China. Biology 10, 955. https://doi.org/10.3390/biology10100955 (2021).
    Google Scholar 
    Kunjaroenruk, J., Koonmanee, S., Singkham, N., Chankaew, S. & Suriharn, K. Genotypic variation and seasonal effects on rice (Oryza sativa L.) grain protein content and yield in tropical savannah environment. J. Agric. Food Res. 20, 101778. https://doi.org/10.1016/j.jafr.2025.101778 (2025).
    Google Scholar 
    Glaz, B. & Kang, M. S. Location contributions determined via GGE biplot analysis of multi-environment sugarcane genotype-performance trials. Crop Sci. 48, 941–950. https://doi.org/10.2135/cropsci2007.06.0315 (2008).
    Google Scholar 
    Oyekunle, M. et al. Assessment of early-maturing maize hybrids and testing sites using GGE biplot analysis. Crop Sci. 57, 2942–2950. https://doi.org/10.2135/cropsci2016.12.1014 (2017).
    Google Scholar 
    Gupta, A. J. et al. Complement test for distinctiveness, uniformity and stability testing of kharif onion (Allium cepa L.) varieties. Genet. Resour. Crop Evol. 69, 2217–2229. https://doi.org/10.1007/s10722-022-01372-z (2022).
    Google Scholar 
    Yan, W. & Kang, M. S. GGE Biplot Analysis: A Graphical Tool for Breeders, Geneticists, and Agronomists 267 (CRC Press, 2003). https://doi.org/10.1201/9781420040371.
    Google Scholar 
    Munawar, M., Hammad, G. & Shahbaz, M. Evaluation of maize (Zea mays L.) hybrids under different environments by GGE biplot analysis. Am.-Euras. J. Agric. Environ. Sci. 13, 1252–1257. https://doi.org/10.5829/idosi.aejaes.2013.13.09.11036 (2013).
    Google Scholar 
    Farshadfar, E., Zali, H. & Mohammadi, R. Evaluation of phenotypic stability in chickpea genotypes using GGE biplot. Ann. Biol. Res. 2(6), 282–292 (2011).
    Google Scholar 
    Rakshit, S. et al. GGE biplot analysis to evaluate genotype, environment and their interactions in sorghum multi-location data. Euphytica 185, 465–479. https://doi.org/10.1007/s10681-012-0648-6 (2012).
    Google Scholar 
    Yang, R. C., Crossa, J., Cornelius, P. L. & Burgueno, J. Biplot analysis of genotype × environment interaction: Proceed with caution. Crop Sci. 49, 1564–1576. https://doi.org/10.2135/cropsci2008.11.0665 (2009).
    Google Scholar 
    Kocaturk, M. et al. GGE biplot analysis of genotype × environment interaction in soybean grown as a second crop. Turk. J. Field Crops 24(2), 145–154. https://doi.org/10.17557/tjfc.615175 (2019).
    Google Scholar 
    Gauch, H. G. & Zobel, R. W. Identifying mega-environments and targeting genotypes. Crop Sci. 37, 311–326. https://doi.org/10.2135/cropsci1997.0011183X003700020002x (1997).
    Google Scholar 
    Mare, M., Manjeru, P., Ncube, B. & Sisito, G. GGE biplot analysis of genotype-by-environment interaction on Sorghum bicolor L. (Moench) in Zimbabwe. Afr. J. Plant Sci. 11, 308–319. https://doi.org/10.5897/AJPS2017.1538 (2017).
    Google Scholar 
    Sousa, M. B. E., Damasceno-Silva, K. J., Rocha, M. D. M., Menezes, J. A. N. & Lima, L. R. L. Genotype by environment interaction in cowpea lines using GGE biplot method. Rev. Caatinga 31, 64–71. https://doi.org/10.1590/1983-21252018v31n108rc (2018).
    Google Scholar 
    Kendal, E. GGE biplot analysis of multi-environment yield trials in barley (Hordeum vulgare L.) cultivars. Ekin J. Crop Breed. Genet. 2(1), 90–99 (2016).
    Google Scholar 
    Yihunie, T. A. & Gesesse, C. A. GGE biplot analysis of genotype by environment interaction in field pea (Pisum sativum L.) genotypes in Northwestern Ethiopia. J. Crop Sci. Biotechnol. 21, 67–74. https://doi.org/10.1007/s12892-017-0099-0 (2018).
    Google Scholar 
    Gungor, H., Cakir, M. F. & Dumlupinar, Z. Evaluation of wheat genotypes: genotype × environment interaction and GGE biplot analysis. Turk. J. Field Crops 27(1), 149–157. https://doi.org/10.17557/tjfc.1081513 (2022).
    Google Scholar 
    Akter, A. et al. GGE biplot analysis for yield stability in multi-environment trials of promising hybrid rice (Oryza sativa L.). Bangladesh Rice J. 19, 1–8. https://doi.org/10.3329/brj.v19i1.25213 (2015).
    Google Scholar 
    Yan, W. & Kang, M. S. GGE Biplot Analysis: A Graphical Tool for Breeders, Geneticists, and Agronomists 1st edn. (CRC Press, 2003). https://doi.org/10.1201/9781420040371.
    Google Scholar 
    Gauch, H. G. A simple protocol for AMMI analysis of yield trials. Crop Sci. 53, 1860–1869. https://doi.org/10.2135/cropsci2013.04.0241 (2013).
    Google Scholar 
    Gauch, H. G. Statistical analysis of yield trials by AMMI and GGE. Crop Sci. 46, 1488–1500 (2006).
    Google Scholar 
    Download referencesAcknowledgementsThe authors sincerely acknowledge networking centers of AINRPOG, ICAR-DOGR for their support and technical inputs. This work was financially supported by Indian Council of Agricultural Research, New Delhi.FundingAuthors gratefully thanks to the networking centers of All India Network Research Project on Onion and Garlic [AINRPOG], ICAR-Directorate of Onion and Garlic Research, Pune for their support and technical inputs. This work was financially supported by Indian Council of Agricultural Research, New Delhi.Author informationAuthors and AffiliationsDepartment of Crop Improvement, ICAR-Directorate of Onion and Garlic Research, Rajgurunagar, Pune, Maharashtra, IndiaAmar Jeet Gupta, Kavya V. Aribenchi, Supriya Kaldate, Hem Raj Bhandari, Pranjali A. Gedam, Yogesh P. Khade, Rajiv B. Kale & Vijay MahajanDepartment of Crop Improvement, ICAR-Indian Agricultural Research Institute, Reginal Sation, Baner, Pune, Maharashtra, IndiaAnil KharAuthorsAmar Jeet GuptaView author publicationsSearch author on:PubMed Google ScholarAnil KharView author publicationsSearch author on:PubMed Google ScholarKavya V. AribenchiView author publicationsSearch author on:PubMed Google ScholarSupriya KaldateView author publicationsSearch author on:PubMed Google ScholarHem Raj BhandariView author publicationsSearch author on:PubMed Google ScholarPranjali A. GedamView author publicationsSearch author on:PubMed Google ScholarYogesh P. KhadeView author publicationsSearch author on:PubMed Google ScholarRajiv B. KaleView author publicationsSearch author on:PubMed Google ScholarVijay MahajanView author publicationsSearch author on:PubMed Google ScholarContributionsAuthor [Amar Jeet Gupta] contributed to the study conception and design. Material preparation and data collection were performed by [Amar Jeet Gupta], [Vijay Mahajan] and [Anil Khar]. The manuscript was prepared by [Amar Jeet Gupta] and [Kavya V. Aribenchi]. Analysis of data and interpretation was performed by [Amar Jeet Gupta] and [Kavya V. Aribenchi] and [Supriya Kaldate]. Manuscript edition and finalization were performed by [Amar Jeet Gupta], [Hem Raj Bhandari], [Pranjali A. Gedam], [Yogesh P. Khade], [Rajiv B. Kale] and [Vijay Mahajan] and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.Corresponding authorCorrespondence to
    Amar Jeet Gupta.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethical approval
    The work is presented in the manuscript with the consent of all authors.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Supplementary Material 2Supplementary Material 3Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleGupta, A.J., Khar, A., Aribenchi, K.V. et al. Assessment of stability and yield performance of onion (Allium cepa L.) genotypes across diverse Indian environments.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-30152-9Download citationReceived: 23 September 2025Accepted: 21 November 2025Published: 24 December 2025DOI: https://doi.org/10.1038/s41598-025-30152-9Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsOnionGGE biplotAMMIStabilityGEIBulb yield More

  • in

    Diversity and persistence of the RNA virome of Philaenus spumarius, the European vector of Xylella fastidiosa

    AbstractPhilaenus spumarius is the primary vector of the quarantine plant pathogen Xylella fastidiosa in Europe and is responsible for the devastating Olive Quick Decline Syndrome outbreak in Southern Italy. Despite its importance, little is known about its natural viral community, which could offer novel and sustainable strategies for vector control. In this three-year study, we conducted the first comprehensive characterization of the viral community of P. spumarius from multiple ecologically diverse European sites, including X. fastidiosa-affected areas in Southern Italy. Deep transcriptomic sequencing of 209 field-collected individuals pooled into 11 RNA-seq libraries revealed the presence of 26 RNA viruses. Our findings revealed a rich and structured viral community in populations from Northern Italy and France, contrasting sharply with the reduced viral diversity observed in populations from Southern Italy, where most individuals were virus-free. Temporal comparisons revealed recurrent virus–host associations over the years, and laboratory rearing provided initial insights into viral persistence and transmission dynamics. Although none of the detected viruses caused overt signs of mortality or sterility, their potential sublethal effects and ecological interactions remain unexplored. This study lays the groundwork for future research on the functional roles of insect-associated viruses and emphasizes their potential for developing sustainable, environmentally friendly approaches to managing vectors and reducing the impact of X. fastidiosa on European agriculture.

    Similar content being viewed by others

    RNA interference protocols for gene silencing in the spittlebug Philaenus spumarius, vector of Xylella fastidiosa

    Article
    Open access
    28 October 2024

    Flight performance and the factors affecting the flight behaviour of Philaenus spumarius the main vector of Xylella fastidiosa in Europe

    Article
    Open access
    02 September 2021

    Expansion of the plant-parasitic nematode RNA viruses: Unprecedented diversity, intron-bearing viruses, and cross-kingdom evolutionary links

    Article
    Open access
    01 December 2025

    Data availability

    All assembled viral genomes and corresponding raw sequencing reads are publicly available in the NCBI GenBank database under the BioProject accession PRJNA1272319.
    ReferencesSanna, F. et al. Ground cover management in Olive groves reduces populations of Philaenus spumarius (Hemiptera: Aphrophoridae), vector of xylella fastidiosa. J. Econ. Entomol. 114 (4), 1716–1721 (2021).
    Google Scholar 
    Dongiovanni, C. et al. Evaluation of efficacy of different insecticides against Philaenus spumarius L., vector of xylella fastidiosa in Olive orchards in Southern Italy, 2015–17. Arthropod Manag Tests. 43 (1), tsy034 (2018).
    Google Scholar 
    Lago, C., Cornara, D., Minutillo, S. A., Moreno, A. & Fereres, A. Feeding behaviour and mortality of Philaenus spumarius exposed to insecticides and their impact on xylella fastidiosa transmission. Pest Manag Sci. 78 (11), 4841–4849 (2022).
    Google Scholar 
    Schneider, K., Barreiro-Hurle, J. & Rodriguez-Cerezo, E. Pesticide reduction amidst food and feed security concerns in Europe. Nat. Food. 4 (9), 746–750 (2023).
    Google Scholar 
    Yousef-Yousef, M. et al. Direct and indirect effects of two endophytic entomopathogenic fungi on survival and feeding behaviour of meadow spittlebug Philaenus spumarius. Biol. Control. 186, 105348 (2023).
    Google Scholar 
    Bodino, N. et al. Activity of natural occurring entomopathogenic fungi on nymphal and adult stages of Philaenus spumarius. J. Invertebr Pathol. 204, 108078 (2024).
    Google Scholar 
    Ganassi, S. et al. Efficacy of entomopathogenic fungi against Philaenus spumarius, the vector of xylella fastidosa. Pest Manag Sci. 80 (9), 4585–4593 (2024).
    Google Scholar 
    Mesmin, X. et al. Ooctonus vulgatus (Hymenoptera, Mymaridae), a potential biocontrol agent to reduce populations of Philaenus spumarius (Hemiptera, Aphrophoridae) the main vector of xylella fastidiosa in Europe. PeerJ 8, e8591 (2020).
    Google Scholar 
    Wagemans, J. et al. Going viral: Virus-Based biological control agents for plant protection. Annu. Rev. Phytopathol. 60, 21–42 (2022).
    Google Scholar 
    Dong, Y. et al. Densovirus infection facilitates plant–virus transmission by an aphid. New. Phytol. 243 (4), 1539–1553 (2024).
    Google Scholar 
    Carvalho, V. L. & Long, M. T. Insect-Specific viruses: an overview and their relationship to arboviruses of concern to humans and animals. Virology 557, 34–43 (2021).
    Google Scholar 
    Roundy, C. M., Azar, S. R., Rossi, S. L., Weaver, S. C. & Vasilakis, N. Insect-Specific viruses: A historical overview and recent developments. Adv. Virus Res. 98, 119–146 (2017).
    Google Scholar 
    Reil, J. B. et al. Transpacific coalescent pathways of coconut rhinoceros beetle biotypes: resistance to biological control catalyses resurgence of an old pest. Mol. Ecol. 27 (22), 4459–4474 (2018).
    Google Scholar 
    Asser-Kaiser, S. et al. Rapid emergence of baculovirus resistance in codling moth due to dominant, sex-linked inheritance. Science 317 (5846), 1916–1918 (2007).
    Google Scholar 
    Cory, J. S. Insect virus transmission: different routes to persistence. Curr. Opin. Insect Sci. 8, 130–135 (2015).
    Google Scholar 
    Sorrell, I., White, A., Pedersen, A. B., Hails, R. S. & Boots, M. The evolution of covert, silent infection as a parasite strategy. Proc. R Soc. B Biol. Sci. 276 (1665), 2217–2226 (2009).
    Google Scholar 
    Cesar, C. S., Miranda, V. H., Silveira, E. R., de Oliveira, T. A. & Cogni, R. Virus infection significantly decreases insect fitness: a meta-analysis [Internet]. 2025 [cited 2025 May 27];2025.01.19.633787. Available from: https://www.biorxiv.org/content/https://doi.org/10.1101/2025.01.19.633787v1Mengual-Martí, A. et al. Impact of Covert infections with an RNA virus on the susceptibility of spodoptera Exigua to natural enemies. BioControl 67 (6), 605–615 (2022).
    Google Scholar 
    Fujiyuki, T. et al. Kakugo virus from brains of aggressive worker honeybees. Adv. Virus Res. 65, 65 (2005).
    Google Scholar 
    Hernández-Pelegrín, L. et al. Covert RNA viruses in medflies differ in their mode of transmission and tissue tropism. J. Virol. 98(6), e00108–e00124 (98AD).
    Google Scholar 
    Witsack, W. & Witsack, W. Experimental and ecological investigations on forms of dormancy in Homoptera-Cicadina (Auchenorrhyncha). 2. on ovarian parapause and obligatory embryonic diapause in Philaenus spumarius (L.) (Aphrophoridae). Zool. Jahrbucher Abt Syst. Okologie Geogr. Tiere. 100 (4), 517–562 (1973).
    Google Scholar 
    Romo, H., Kenney, J. L., Blitvich, B. J. & Brault, A. C. Restriction of Zika virus infection and transmission in Aedes aegypti mediated by an insect-specific flavivirus. Emerg. Microbes Infect. 7 (1), 181 (2018).
    Google Scholar 
    Kenney, J. L., Solberg, O. D., Langevin, S. A. & Brault, A. C. Characterization of a novel insect-specific flavivirus from brazil: potential for Inhibition of infection of arthropod cells with medically important flaviviruses. J. Gen. Virol. 95 (Pt 12), 2796–2808 (2014).
    Google Scholar 
    Hall-Mendelin, S. et al. The insect-specific palm creek virus modulates West nile virus infection in and transmission by Australian mosquitoes. Parasit. Vectors. 9 (1), 414 (2016).
    Google Scholar 
    Carvalho, V. L. et al. Negevirus Piura Suppresses Zika Virus Replication Mosq. Cells Viruses ;16(3):350. (2024).
    Google Scholar 
    Shi, M. et al. Redefining the invertebrate RNA virosphere. Nature 540 (7634), 539–543 (2016).
    Google Scholar 
    Wu, H. et al. Abundant and Diverse RNA Viruses in Insects Revealed by RNA-Seq Analysis: Ecological and Evolutionary Implications. mSystems. ;5(4): (2020). https://doi.org/10.1128/msystems.00039Sun, K., Fu, K., Hu, T., Shentu, X. & Yu, X. Leveraging insect viruses and genetic manipulation for sustainable agricultural pest control. Pest Manag Sci. 80 (6), 2515–2527 (2024).
    Google Scholar 
    Bichaud, L. et al. Arthropods as a source of new RNA viruses. Microb. Pathog. 77, 136–141 (2014).
    Google Scholar 
    Ros, V. I. D. et al. Viral diseases of insects [Internet]. In: Rowley AF, Coates CJ, Whitten MW, editors. Invertebrate Pathology. Oxford University Press; [cited 2025 Nov 28]. page 0. (2022). Available from: https://doi.org/10.1093/oso/9780198853756.003.0010Haas, B. J. et al. De Novo transcript sequence reconstruction from RNA-seq using the trinity platform for reference generation and analysis. Nat. Protoc. 8 (8), 1494–1512 (2013).
    Google Scholar 
    Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17 (1), 10–12 (2011).
    Google Scholar 
    Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10 (3), R25 (2009).
    Google Scholar 
    Kuhn, J. H. et al. Promotion of order bunyavirales to class bunyaviricetes to accommodate a rapidly increasing number of related Polyploviricotine viruses. J. Virol. 98 (10), e01069–e01024 (2024).
    Google Scholar 
    Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30 (4), 772–780 (2013).
    Google Scholar 
    Minh, B. Q. et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37 (5), 1530–1534 (2020).
    Google Scholar 
    Letunic, I. & Bork, P. Interactive tree of life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 49 (W1), W293–W296 (2021).
    Google Scholar 
    Obbard, D. J., Shi, M., Roberts, K. E., Longdon, B. & Dennis, A. B. A new lineage of segmented RNA viruses infecting animals. Virus Evol. 6 (1), vez061 (2020).
    Google Scholar 
    Zhang, X., Wang, N., Wang, Z. & Liu, Q. The discovery of segmented flaviviruses: implications for viral emergence. Curr. Opin. Virol. 40, 11–18 (2020).
    Google Scholar 
    Valle, C., Parry, R. H., Coutard, B. & Colmant, A. M. G. Discovery of additional genomic segments reveals the fluidity of jingmenvirus genomic organization. Virus Evol. 11 (1), veaf023 (2025).
    Google Scholar 
    Bodino, N. et al. Spittlebugs of mediterranean Olive groves: Host-Plant exploitation throughout the year. Insects 11 (2), 130 (2020).
    Google Scholar 
    Bodino, N. et al. Phenology, seasonal abundance and stage-structure of spittlebug (Hemiptera: Aphrophoridae) populations in Olive groves in Italy. Sci. Rep. 9 (1), 17725 (2019).
    Google Scholar 
    Biedermann, R. Patch occupancy of two hemipterans sharing a common host plant. J. Biogeogr. 31 (7), 1179–1184 (2004).
    Google Scholar 
    Paraskevopoulou, S. et al. Viromics of extant insect orders unveil the evolution of the flavi-like superfamily. Virus Evol. 7 (1), veab030 (2021).
    Google Scholar 
    Dolja, V. V. & Koonin, E. V. Metagenomics reshapes the concepts of RNA virus evolution by revealing extensive horizontal virus transfer. Virus Res. 244, 36–52 (2018).
    Google Scholar 
    WolfYI et al. Origins and evolution of the global RNA Virome. mBio 9 (6), 101128mbio02329–101128mbio02318 (2018).
    Google Scholar 
    Chiapello, M. et al. Complexity and local specificity of the virome associated with tospovirus-transmitting thrips species. J. Virol. https://doi.org/10.1128/jvi.00597 (2021).
    Google Scholar 
    Abbo, S. R. et al. The invasive Asian Bush mosquito Aedes japonicus found in the Netherlands can experimentally transmit Zika virus and Usutu virus. PLoS Negl. Trop. Dis. 14 (4), e0008217 (2020).
    Google Scholar 
    Shi, C. et al. Stable distinct core eukaryotic Viromes in different mosquito species from Guadeloupe, using single mosquito viral metagenomics. Microbiome 7 (1), 121 (2019).
    Google Scholar 
    Hollingsworth, B. D., Grubaugh, N. D., Lazzaro, B. P. & Murdock, C. C. Leveraging insect-specific viruses to elucidate mosquito population structure and dynamics. PLOS Pathog. 19 (8), e1011588 (2023).
    Google Scholar 
    Moonen, J. P., Schinkel, M., van der Most, T., Miesen, P. & van Rij, R. P. Composition and global distribution of the mosquito virome – A comprehensive database of insect-specific viruses. One Health Amst Neth. 16, 100490 (2023).
    Google Scholar 
    Zhou, M. et al. Horizontal and vertical transmission of wild-type and Recombinant Helicoverpa armigera single-nucleocapsid nucleopolyhedrovirus. J. Invertebr Pathol. 89 (2), 165–175 (2005).
    Google Scholar 
    Morris, M. M., Frixione, N. J., Burkert, A. C., Dinsdale, E. A. & Vannette, R. L. Microbial abundance, composition, and function in nectar are shaped by flower visitor identity. FEMS Microbiol. Ecol. 96 (3), fiaa003 (2020).
    Google Scholar 
    Ottati, S. et al. Biological characterization of Euscelidius variegatus iflavirus 1. J. Invertebr Pathol. 173, 107370 (2020).
    Google Scholar 
    Longdon, B. & Jiggins, F. M. Vertically transmitted viral endosymbionts of insects: do Sigma viruses walk alone? Proc. Biol. Sci. 279 (1744), 3889–3898 (2012).
    Google Scholar 
    Touret, F., Guiguen, F., Greenland, T. & Terzian, C. In between: Gypsy in drosophila melanogaster reveals new insights into endogenous retrovirus evolution. Viruses 6 (12), 4914–4925 (2014).
    Google Scholar 
    Mao, Q. et al. Viral pathogens hitchhike with insect sperm for paternal transmission. Nat. Commun. 10 (1), 955 (2019).
    Google Scholar 
    Ros, V. I. D. et al. Viral diseases of insects [Internet]. In: Rowley AF, Coates CJ, Whitten MW, editors. Invertebrate Pathology. Oxford University Press; [cited 2025 July 14]. page 0. (2022). Available from: https://doi.org/10.1093/oso/9780198853756.003.0010Carballo, A., Williams, T., Murillo, R. & Caballero, P. Iflavirus Covert infection increases susceptibility to nucleopolyhedrovirus disease in spodoptera Exigua. Viruses 12 (5), 509 (2020).
    Google Scholar 
    Download referencesAcknowledgementsWe would like to thank Massimo Turina for the critical revision of the text and Luca Bucci for the technical assistance.FundingThis work was supported by the European Union’s Horizon Europe, BeXyL project [grant number 101060593] and Italian Ministry of Agriculture, MASAF, SOS project “Sviluppo di Strategie di controllo sostenibili di Philaenus spumarius ed interferenza con la trasmissione di Xylella fastidiosa”.Author informationAuthors and AffiliationsInstitute for Sustainable Plant Protection, IPSP-CNR, Strada delle Cacce 73, Torino, ItalySara Ottati, Luciana Galetto, Cristina Marzachí & Simona AbbáDepartment of Agriculture, Forest and Food Sciences, University of Torino, Largo Paolo Braccini 2, Grugliasco, ItalySara Ottati, Francesco Volpe, Nicola Bodino & Domenico BoscoAuthorsSara OttatiView author publicationsSearch author on:PubMed Google ScholarLuciana GalettoView author publicationsSearch author on:PubMed Google ScholarFrancesco VolpeView author publicationsSearch author on:PubMed Google ScholarNicola BodinoView author publicationsSearch author on:PubMed Google ScholarCristina MarzachíView author publicationsSearch author on:PubMed Google ScholarDomenico BoscoView author publicationsSearch author on:PubMed Google ScholarSimona AbbáView author publicationsSearch author on:PubMed Google ScholarContributionsConceptualization, D.B., C.M. and S.O.; methodology, S.O. and S.A.; formal analysis, S.O.; investigation, S.O., S.A. and L.G.; software, S.O. and S.A.; writing—original draft preparation, S.O. and S.A.; writing—review and editing, S.O., S.A., L.G., M.C., C.M., F.V., N.B. and D.B.; field sampling and insect rearing, S.O., F.V. and N.B.; visualization, S.O. and S.A.; supervision, S.A. and D.B.; resources, D.B. and C.M.; project administration, D.B and C.M. All authors have read and agreed to the published version of the manuscript.Corresponding authorCorrespondence to
    Domenico Bosco.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleOttati, S., Galetto, L., Volpe, F. et al. Diversity and persistence of the RNA virome of Philaenus spumarius, the European vector of Xylella fastidiosa.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-32597-4Download citationReceived: 15 September 2025Accepted: 11 December 2025Published: 24 December 2025DOI: https://doi.org/10.1038/s41598-025-32597-4Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    Keywords
    Philaenus spumarius

    Xylella fastidiosa
    Insect-specific-virusTransmission routesPlant pestBiocontrol More

  • in

    Spatial and temporal assessment of soil degradation risk in Europe

    Abstract

    Soil degradation threatens agricultural productivity and ecosystem resilience across Europe, yet spatially consistent assessments of its intensity and drivers remain limited. In this study, we used Soil Degradation Proxy (SDP), that integrates four key indicators of soil degradation, including erosion rate, soil pH, electrical conductivity, and organic carbon content, to quantify soil degradation risk. Using over 38,000 LUCAS topsoil observations and a machine learning model trained on climate, land cover, topographic, soil parent material properties, and spectral variables, we map annual SDP values between years 2000 to 2022 across Europe. Results show soil degradation risk is highest in southern Europe, especially in intensively managed and sparsely vegetated landscapes. Over the past two decades, approximately 7.1% of land area across the EU and the UK has experienced increasing degradation risk (most notably across Eastern Europe), with rainfed croplands emerging as the most affected land cover type. Land cover is the most influential driver, modulating effects of climatic variables such as precipitation and temperature on SDP. This data-driven framework provides a consistent and scalable approach for monitoring soil degradation risk and offers actionable insights to support targeted conservation and EU-wide policy implementation.

    Data availability

    The required data of the study explained in Sect. Results were obtained from the following sources: The LUCAS observations are available at https://esdac.jrc.ec.europa.eu/content/lucas2015-topsoil-data, elevation data are obtained from https:/doi.org/10.5270/ESA-c5d3d65, MODIS observations are retrieved from https://doi.org/10.5067/MODIS/MOD11A2.061; https://doi.org/10.5067/MODIS/MOD13A2.061; and https://doi.org/10.5067/MODIS/MOD09GA.061, soil parent material properties were downloaded from SoilGrids https://doi.org/10.17027/isric-soilgrids.713396fa-1687-11ea-a7c0-a0481ca9e724, land cover datasets are taken from Copernicus Global Land Service https://doi.org/10.24381/cds.006f2c9a, and lithology maps are obtained from https://doi.org/10.5281/zenodo.12607973.
    ReferencesJohnson, D. L. et al. Meanings of environmental terms. J. Environ. Qual. 26, 581–589 (1997).
    Google Scholar 
    Panagos, P., Borrelli, P., Jones, A. & Robinson, D. A. A 1 billion Euro mission: A soil deal for Europe. Eur J. Soil. Sci 75, e13466 (2024).Feddema, J. J. & Freire, S. Soil degradation, global warming and climate impacts. Clim. Res. 17, 209–216 (2001).
    Google Scholar 
    Gregory, A. S. et al. A review of the impacts of degradation threats on soil properties in the UK. Soil. Use Manag. 31, 1–15 (2015).
    Google Scholar 
    Shokri, N. et al. Rethinking Global Soil Degradation: Drivers, Impacts, and Solutions. Reviews of Geophysics 63, e2025RG000883 (2025). EU Commission. Thematic Strategy for Soil Protection. Communication from the Commission to the Council, the European Parliament, the European Economic and Social Committee and the Committee of the Regions (2006). https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX:52006DC0231Ferreira, C. S. S., Seifollahi-Aghmiuni, S., Destouni, G., Ghajarnia, N. & Kalantari, Z. Soil degradation in the European Mediterranean region: Processes, status and consequences. Sci. Total Environ. 805, 150106 (2022).
    Google Scholar 
    Dubovyk, O. The role of remote sensing in land degradation assessments: opportunities and challenges. Eur. J. Remote Sens. 50, 601–613 (2017).
    Google Scholar 
    Yang, Y. et al. Climate change exacerbates the environmental impacts of agriculture. Science. 385, eadn3747 (2024).Panagos, P. et al. Projections of soil loss by water erosion in Europe by 2050. Environ. Sci. Policy. 124, 380–392 (2021).
    Google Scholar 
    Borrelli, P. et al. Land use and climate change impacts on global soil erosion by water (2015–2070). Proc. Natl. Acad. Sci. 117, 21994–22001 (2020).
    Google Scholar 
    Rinot, O., Levy, G. J., Steinberger, Y., Svoray, T. & Eshel, G. Soil health assessment: A critical review of current methodologies and a proposed new approach. Sci. Total Environ. 648, 1484–1491 (2019).
    Google Scholar 
    Lehmann, J., Bossio, D. A. & Kögel-Knabner, I. & others. The concept and future prospects of soil health. Nat Rev Earth Environ 1, 544–553 (2020). Acton, D. F. & Gregorich, L. J. The health of our soils: toward sustainable agriculture in Canada. (1995). Armenise, E., Redmile-Gordon, M. A., Stellacci, A. M., Ciccarese, A. & Rubino, P. Developing a soil quality index to compare soil fitness for agricultural use under different managements in the Mediterranean environment. Soil. Tillage Res. 130, 91–98 (2013).
    Google Scholar 
    Moebius-Clune, B. N. Comprehensive Assessment of Soil Health: The Cornell Framework ManualCornell University (2016). Feeney, C. J. et al. Development of soil health benchmarks for managed and semi-natural landscapes. Sci. Total Environ. 886, 163973 (2023).
    Google Scholar 
    Prăvălie, R. et al. A unifying modelling of multiple land degradation pathways in Europe. Nat. Commun. 15, 3862 (2024).
    Google Scholar 
    Gianoli, F., Weynants, M. & Cherlet, M. Land degradation in the European Union—Where does the evidence converge? Land. Degrad. Dev. 34, 2256–2275 (2023).
    Google Scholar 
    Prăvălie, R. et al. Arable lands under the pressure of multiple land degradation processes. A global perspective. Environ. Res. 194, 110697 (2021).
    Google Scholar 
    Yang, L. et al. Integrated assessments of land degradation on the Qinghai-Tibet plateau. Ecol. Indic. 147, 109945 (2023).
    Google Scholar 
    Cherif, I., Kolintziki, E. & Alexandridis, T. K. Monitoring of land degradation in Greece and Tunisia using Trends.Earth with a focus on cereal croplands. Remote Sens. (Basel). 15, 1766 (2023).
    Google Scholar 
    Lu, A. et al. Fuzzy logic modeling of land degradation in a loess plateau Watershed, China. Remote Sens. (Basel). 14, 4779 (2022).
    Google Scholar 
    Panagos, P. et al. How the EU soil observatory is providing solid science for healthy soils. Eur. J. Soil. Sci. 75, e13507 (2024).
    Google Scholar 
    Orgiazzi, A., Ballabio, C., Panagos, P. & Jones, A. Fernández-Ugalde, O. LUCAS Soil, the largest expandable soil dataset for Europe: A review. Eur. J. Soil. Sci. 69, 140–153 (2018).
    Google Scholar 
    Hassani, A., Azapagic, A. & Shokri, N. Global predictions of primary soil salinization under changing climate in the 21st century. Nature Communications 2021 (12:1 12), 1–17 (2021). Balkovič, J. et al. Impacts and uncertainties of + 2°C of climate change and soil degradation on European crop calorie supply. Earths Future. 6, 373–395 (2018).
    Google Scholar 
    Lal, R. Climate change and soil degradation mitigation by sustainable management of soils and other natural resources. Agricultural Res. 1, 199–212 (2012).
    Google Scholar 
    Borrelli, P. et al. An assessment of the global impact of 21st century land use change on soil erosion. Nat. Commun. 8, 1–13 (2017).
    Google Scholar 
    Montanarella, L. & Panagos, P. The relevance of sustainable soil management within the European green deal. Land. Use Policy. 100, 104950 (2021).
    Google Scholar 
    Abdalla, K., Mutema, M. & Hill, T. Soil and organic carbon losses from varying land uses: a global meta-analysis. Geographical Res. 58, 167–185 (2020).
    Google Scholar 
    Baumhardt, R., Stewart, B. & Sainju, U. North American soil degradation: Processes, Practices, and mitigating strategies. Sustainability 7, 2936–2960 (2015).
    Google Scholar 
    Katayama, A. et al. Short communication: concentrated impacts by tree canopy drips – hotspots of soil erosion in forests. Earth Surf. Dyn. 11, 1275–1282 (2023).
    Google Scholar 
    Soares, P. R., Harrison, M. T., Kalantari, Z., Zhao, W. & Ferreira, C. S. S. Drought effects on soil organic carbon under different agricultural systems. Environ. Res. Commun. 5, 112001 (2023).
    Google Scholar 
    Dang, P. et al. Long-term climate warming weakens positive plant biomass responses globally. Proceedings of the National Academy of Sciences 122, e2420379122 (2025). Antoni, V. et al. The state of the soils in France in 2011-A synthesis. Groupement d’Intêret Scientifique sur les Sols, 24 ISBN 978-2-7380-1296-8. (2013).Pereira, P. et al. Agricultural soil degradation in Estonia, Latvia and Lithuania. in Impact of Agriculture on Soil Degradation II: A European Perspective (ed. Pereira, P., Muñoz-Rojas, M., Bogunovic, I., Zhao, W) 59–86 (Springer, 2023).Wadoux, A. M. J. C. et al. On soil districts. Geoderma 452, 117065 (2024).
    Google Scholar 
    Harvey, L. & Green, D. Defining quality. Assess. Eval High. Educ. 18, 9–34 (1993).
    Google Scholar 
    Svoray, T., Hassid, I., Atkinson, P. M., Moebius-Clune, B. N. & van Es, H. M. Mapping soil health over large agriculturally important areas. Soil Sci. Soc. Am. J. 79, 1420–1434 (2015).
    Google Scholar 
    Campbell, G. A. et al. Continental scale soil monitoring: A proposed Multi-Scale framing of soil quality. Eur. J. Soil. Sci. 76, e70174 (2025).
    Google Scholar 
    Borrelli, P. et al. An assessment of the global impact of 21st century land use change on soil erosion. Nature Communications 2017 (8:1 8), 1–13 (2017). Alewell, C. et al. Global phosphorus shortage will be aggravated by soil erosion. Nature Communications 2020 (11:1 11), 1–12 (2020). Kalantari, Z. et al. Assessing flood probability for transportation infrastructure based on catchment characteristics, sediment connectivity and remotely sensed soil moisture. Sci. Total Environ. 661, 393–406 (2019).
    Google Scholar 
    Panagos, P. et al. Cost of agricultural productivity loss due to soil erosion in the European union: from direct cost evaluation approaches to the use of macroeconomic models. Land. Degrad. Dev. 29, 471–484 (2018).
    Google Scholar 
    Singh, A. Soil salinization management for sustainable development: A review. J. Environ. Manage. 277, 111383 (2021).
    Google Scholar 
    ZHANG, W., wen, W. A. N. G. & XUE, C. WANG, L. jie. Effects of salinity on the soil microbial community and soil fertility. J. Integr. Agric. 18, 1360–1368 (2019).
    Google Scholar 
    Majeed, A., Muhammad, Z. & Salinity A Major Agricultural Problem—Causes, Impacts on Crop Productivity and Management Strategies. In: Plant Abiotic Stress Tolerance 83–99 (Springer International Publishing, Cham, 2019). https://doi.org/10.1007/978-3-030-06118-0_3Shrivastava, P. & Kumar, R. Soil salinity: A serious environmental issue and plant growth promoting bacteria as one of the tools for its alleviation. Saudi J. Biol. Sci. 22, 123 (2014).
    Google Scholar 
    Aciego Pietri, J. C. & Brookes, P. C. Relationships between soil pH and microbial properties in a UK arable soil. Soil. Biol. Biochem. 40, 1856–1861 (2008).
    Google Scholar 
    Rousk, J. et al. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 2010. 4:10 (4), 1340–1351 (2010).
    Google Scholar 
    Lebron, I. et al. Patterns and thresholds for soil pH across Europe in relation to soil health and degradation. Catena (Amst). 260, 109454 (2025).
    Google Scholar 
    Rengel, Z. Soil & pH Soil Health and Climate Change. 69–85 (2011). https://doi.org/10.1007/978-3-642-20256-8_4Obalum, S. E., Chibuike, G. U., Peth, S. & Ouyang, Y. Soil organic matter as sole indicator of soil degradation. Environ. Monit. Assess. 189, 176 (2017).
    Google Scholar 
    Lal, R. Soil erosion impact on agronomic productivity and environment quality. CRC Crit. Rev. Plant. Sci. 17, 319–464 (1998).
    Google Scholar 
    Fageria, N. K. Role of soil organic matter in maintaining sustainability of cropping systems. Commun. Soil. Sci. Plant. Anal. 43, 2063–2113 (2012).
    Google Scholar 
    Plante, A. F. & Parton, W. J. The dynamics of soil & organic matter and nutrient cycling. Soil. Microbiol. Ecol. Biochemistry: Third Ed. 433–467 https://doi.org/10.1016/B978-0-08-047514-1.50020-2 . (2007).Taylor, K. C. et al. The role of soil organic matter in sustaining soil fertility. Nature 1994 (371:6500), 783–785 (1994). Ghorbani, R., Wilcockson, S., Koocheki, A. & Leifert, C. Soil management for sustainable crop disease control: A review. Environ. Chem. Lett. 6, 149–162 (2008).
    Google Scholar 
    Haddaway, N. R. et al. How does tillage intensity affect soil organic carbon? A systematic review. Environ. Evid. 6, 1–48 (2017).
    Google Scholar 
    Goulding, K. W. T. Soil acidification and the importance of liming agricultural soils with particular reference to the United Kingdom. Soil. Use Manag. 32, 390–399 (2016).
    Google Scholar 
    Läuchli, A. & Grattan, S. R. Soil pH extremes. in Plant Stress Physiology 194–209 (CABI, UK, doi:https://doi.org/10.1079/9781845939953.0194. (2012).
    Google Scholar 
    Oshunsanya, S. O. Introductory chapter: Relevance of soil pH to agriculture. In: Soil PH for nutrient availability and crop performance (IntechOpen, 2018). Slessarev, E. W. et al. Water balance creates a threshold in soil pH at the global scale. Nature 540, 567–569 (2016).
    Google Scholar 
    Ballabio, C. et al. Mapping monthly rainfall erosivity in Europe. Sci. Total Environ. 579, 1298–1315 (2017).
    Google Scholar 
    Panagos, P., Borrelli, P. & Meusburger, K. A new European slope length and steepness factor (LS-Factor) for modeling soil erosion by water. Geosci. 2015. 5, Pages 117–126 (5), 117–126 (2015).
    Google Scholar 
    Panagos, P., Borrelli, P. & Robinson, D. A. Tackling soil loss across Europe. Nature 526, 195–195 (2015).
    Google Scholar 
    Panagos, P. et al. Modelling the effect of support practices (P-factor) on the reduction of soil erosion by water at European scale. Environ. Sci. Policy. 51, 23–34 (2015).
    Google Scholar 
    Panagos, P. et al. Estimating the soil erosion cover-management factor at the European scale. Land. Use Policy. 48, 38–50 (2015).
    Google Scholar 
    Panagos, P. et al. The new assessment of soil loss by water erosion in Europe. Environ. Sci. Policy. 54, 438–447 (2015).
    Google Scholar 
    Copernicus, D. E. M. Global and European Digital Elevation Model | Copernicus Data Space Ecosystem. https://doi.org/10.5270/ESA-c5d3d65Copernicus Climate Change Service (C3S) Climate Data Store (CDS). Copernicus Climate Change Service, Climate Data Store. : Land cover classification gridded maps from 1992 to present derived from satellite observation. (2019). https://doi.org/10.24381/cds.006f2c9a (2019).Isik, S., Minarik, R. & Hengl, T. Continental Europe surface lithology based on EGDI / OneGeology map at 1:1 M scale. Preprint at (2024). https://doi.org/10.5281/zenodo.12607973Muñoz-Sabater, J. et al. ERA5-Land: A state-of-the-art global reanalysis dataset for land applications. Earth Syst. Sci. Data. 13, 4349–4383 (2021).
    Google Scholar 
    Didan, K. MODIS/Terra Vegetation Indices 16-Day L3 Global 1km SIN Grid V061. (2021). https://doi.org/10.5067/MODIS/MOD13A2.061Vermote, E. & Wolfe, R. MODIS/Terra surface reflectance daily L2G global 1km and 500m SIN Grid V061. https://doi.org/10.5067/MODIS/MOD09GA.061Beeson, P. C., Daughtry, C. S. T. & Wallander, S. A. Estimates of conservation tillage practices using Landsat archive. Remote Sens. 2020. 12, Page 2665 (12), 2665 (2020).
    Google Scholar 
    Poggio, L. et al. SoilGrids 2.0: producing soil information for the Globe with quantified Spatial uncertainty. SOIL 7, 217–240 (2021).
    Google Scholar 
    Hengl, T. et al. SoilGrids250m: global gridded soil information based on machine learning. PLoS One. 12, e0169748 (2017).
    Google Scholar 
    Apley, D. W. & Zhu, J. Visualizing the effects of predictor variables in black box supervised learning models. J. R Stat. Soc. Ser. B Stat. Methodol. 82, 1059–1086 (2020).
    Google Scholar 
    Gorelick, N. et al. Google Earth engine: Planetary-scale Geospatial analysis for everyone. Remote Sens. Environ. 202, 18–27 (2017).
    Google Scholar 
    Liaw, A. & Wiener, M. Classification and regression by randomforest. R News. 2, 18–22 (2002).
    Google Scholar 
    Kahle, D. & Wickham, H. ggmap: Spatial Visualization with ggplot2. The R Journal. 5, 144–161 (2013).Kuhn, M. Caret package. J. Stat. Softw. 28, 1–26 (2008).
    Google Scholar 
    Hofert, M. et al. Package ‘copula’. URL (2014). http://ie.archive.ubuntu.com/disk1/disk1/cran. r-project. org/web/packages/copula/copula. pdfSouth, A. rworldmap: a new R package for mapping global data. (2011).Hijmans, R. J. et al. Package ‘terra’.https://github.com/rspatial/terra (2022). R Core Team. R: A language and environment for statistical computing. (2023).Download referencesAcknowledgementsThe authors thank the European Commission’s Joint Research Centre for access to LUCAS datasets, the Copernicus programme for providing land cover data, the European Space Agency for DEM data, the NASA MODIS team for vegetation and tillage indices, ISRIC – SoilGrids for global soil property data, the ECMWF for ERA5-Land climate data, and OpenGeoHub for providing the lithology dataset, derived from the EGDI / OneGeology 1:1 M scale map.FundingOpen Access funding enabled and organized by Projekt DEAL. This research is part of the project AI4SoilHealth (Accelerating collection and use of soil health information using AI technology to support the Soil Deal for Europe and EU Soil Observatory) to support the Soil Deal for Europe and EU Soil funded Horizon Europe (Grant No. 101086179). P.B. was funded by the Swiss State Secretariat for Education, Research and Innovation (SERI), grant agreement no. 101086179 (Horizon Europe name AI4SoilHealth).Author informationAuthors and AffiliationsInstitute of Geo-Hydroinformatics, Hamburg University of Technology, Hamburg, GermanyMehdi H. Afshar, Milad Aminzadeh & Nima ShokriUnited Nations University Hub on Engineering to Face Climate Change at the Hamburg University of Technology, United Nations University Institute for Water, Environment and Health (UNU-INWEH), Hamburg, GermanyMehdi H. Afshar, Milad Aminzadeh & Nima ShokriThe Climate and Environmental Research Institute NILU, P.O. Box 100, Kjeller, 2027, NorwayAmirhossein HassaniDepartment of Environmental Sciences, University of Basel, Basel, 4056, SwitzerlandPasquale BorrelliEuropean Commission, Joint Research Centre (JRC), Ispra, ItalyPanos PanagosCentre for Ecology & Hydrology, Bangor, UKDavid A. RobinsonDepartment of Civil & Environmental Engineering, University of Nevada, Reno, NV, USADani OrEnvironmental Systems Science, ETH Zurich, Zurich, SwitzerlandDani OrAuthorsMehdi H. AfsharView author publicationsSearch author on:PubMed Google ScholarAmirhossein HassaniView author publicationsSearch author on:PubMed Google ScholarMilad AminzadehView author publicationsSearch author on:PubMed Google ScholarPasquale BorrelliView author publicationsSearch author on:PubMed Google ScholarPanos PanagosView author publicationsSearch author on:PubMed Google ScholarDavid A. RobinsonView author publicationsSearch author on:PubMed Google ScholarDani OrView author publicationsSearch author on:PubMed Google ScholarNima ShokriView author publicationsSearch author on:PubMed Google ScholarContributionsConceptualization: All authors, Methodology: MHA, AH, Validation: All authors, Formal analysis: MHA, Investigation: All authors, Data Curation: MHA, AH, Writing – Original Draft: MHA, Writing – Review & Editing: All authors, Visualization: MHA, Supervision: NS, Funding acquisition: NS.Corresponding authorsCorrespondence to
    Mehdi H. Afshar or Nima Shokri.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary InformationBelow is the link to the electronic supplementary material.Supplementary Material 1Supplementary Material 2Rights and permissions
    Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
    Reprints and permissionsAbout this articleCite this articleAfshar, M.H., Hassani, A., Aminzadeh, M. et al. Spatial and temporal assessment of soil degradation risk in Europe.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-33318-7Download citationReceived: 06 August 2025Accepted: 17 December 2025Published: 24 December 2025DOI: https://doi.org/10.1038/s41598-025-33318-7Share this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative
    KeywordsSoil monitoringMachine learningLand coverClimate variabilitySoil health indicatorEnvironmental policy More

  • in

    A high-resolution global leaf chlorophyll content product using the Sentinel-2 data

    AbstractLeaf chlorophyll content (LCC) serves as a critical indicator for quantifying photosynthetic carbon assimilation, providing fundamental data for terrestrial carbon cycle estimation. In the past five years, some global LCC remote sensing products have been generated, but their resolution ranges from 300 m to 500 m. This study employed an empirical relationship method based on the Chlorophyll Sensitive Index (CSI) to produce the Multi-source data Synergized Quantitative Global LCC product (MuSyQ Global LCC) with a resolution from 100 m to 10 m using the Google Earth Engine (GEE) platform. Validation results demonstrate that the 10m-resolution MuSyQ Global LCC product has an RMSE of 13.69 μg/cm2, R2 of 0.37, and the RMSE is between 11.28 μg/cm2 and 15.22 μg/cm2 for different vegetation types. Its finer resolution reveals more spatial details compared with the existing global products. When upscaled to 500 m, it demonstrates high consistency with the MODIS LCC product, and MuSyQ Global LCC (RMSE = 14.16 μg/cm2) exhibits higher accuracy than MODIS LCC (RMSE = 14.74 μg/cm2).

    Similar content being viewed by others

    Selection of optimal spectral features for leaf chlorophyll content estimation

    Article
    Open access
    27 October 2024

    Mind the leaf anatomy while taking ground truth with portable chlorophyll meters

    Article
    Open access
    13 January 2025

    An improved spatially downscaled solar-induced chlorophyll fluorescence dataset from the TROPOMI product

    Article
    Open access
    22 January 2025

    Background & SummaryChlorophyll is an essential pigment in green plants’ photosynthesis that harvests solar radiation and absorbs carbon dioxide. Leaf chlorophyll content (LCC) indicates the maximum carboxylation rate (Vcmax)1 and can then be used to calculate the primary productivity of plants2. LCC also indicates light, temperature, water stress, pests, and diseases. Therefore, the accurate large-scale LCC can improve the performance of the terrestrial global carbon cycle model3,4 and the ability of ecosystem monitoring. Remote sensing methods, taking advantage of chlorophyll’s varied absorption and scattering properties in different bands, make it the only applicable approach to retrieve LCC at the continental or global scale.In the past decades, some global LCC products have been generated. MERIS LCC product is the first global LCC whose spatial resolution is 300 m and temporal resolution is 7 days5. Leaf-level radiative transfer model PROSPECT combined with the 4-Scale model (for woody vegetation) and SAIL model (for non-woody vegetation) were used to construct look-up table (LUT) and derive LCC from the MERIS data. MODIS LCC is the product generated by MODIS data from a VI matrix method6. The spatial resolution is 500 m, and the temporal resolution is 8 days. GLCC products were derived from ENVISAT MERIS and Sentinel-3 OLCI with a spatial resolution of 500 m and a temporal resolution of 7 days7. LUTs constructed from the PROSPECT-D + 4-Scale model (for heterogeneous vegetation) and PROSPECT-D + 4SAIL model (for homogenous vegetation) were used to derive global LCC. Another global LCC product is the GLOBMAP MERIS LCC8. Based on the RTM simulations, a neural network was constructed and derived from the ENVISAT MERIS LCC with a resolution of 300 m/7 days from 2003 to 2012. The current global LCC product’s spatial resolution is from 300 m to 500 m. A recent study compared the performance of these LCC products in China and showed that the RMSE ranged from 21.0 μg/cm2 to 32.3 μg/cm2  9, indicating the accuracy still requires systematic improvement. In the scale of 300–500 m, a large proportion of vegetation areas should be in the mixed pixels10, and the mixed-pixel effect brings great uncertainties to the inversion of vegetation parameters11. Enhancing the spatial resolution becomes a practical way to improve the accuracy of the current product12, which allows researchers and analysts to better monitor vegetation status13 and predict the crop yield14. Meanwhile, the higher-resolution product is a more effective reference for decision-making in fine precision agriculture15 and grazing a more reasonable input for global and regional ecosystem models associated with carbon cycle modeling16.The only published large-scale and high-resolution LCC product is the Multi-source data Synergized Quantitative remote sensing production system LCC (MuSyQ LCC17,18). Using Sentinel-2 MSI reflectance and the Chlorophyll Sensitive Index (CSI)-based empirical regression method, the resolution of the LCC product was improved to 30 m/10 days19. Previous study also suggests the CSI-based algorithm is stable and suitable for generating the large-scale LCC product when applied to Gaofen-6 images20. Validation suggests the MuSyQ LCC has the highest accuracy, demonstrating high overall spatial consistency with the MODIS LCC over China9. However, the MuSyQ LCC product only covers China from 2019 to 2020 without high-resolution global LCC information.The objectives of this study are twofold: (1) To generate a high-resolution global LCC product by establishing empirical relationships between CSI and LCC through radiative transfer model, with the 100m-resolution product archived in Science Data Bank and higher-resolution versions accessible via provided code; (2) To validate the multi-resolution LCC products through direct comparison with ground measurements and indirect evaluation against MODIS LCC products.MethodsInput dataSentinel-2 Multispectral Instrument (MSI) images were used to generate the MuSyQ Global LCC product. The MSI onboard Sentinel-2 has 13 bands, including red-edge bands sensitive to the LCC variation. The spatial resolution of Sentinel-2 is 10 m for visible and near-infrared (NIR) bands and 20 m for the red-edge bands. The Sentinel-2 MSI level 2 (L2A) land surface reflectance product, pre-processed with radiometric calibration, geometric and atmospheric correction, is an ideal dataset for calculating the CSI index and retrieving the LCC. The Sentinel-2 MSI L2A dataset is available on both the official website (https://dataspace.copernicus.eu/) and the Google Earth Engine (GEE) platform (https://developers.google.com/earth-engine/datasets/catalog/COPERNICUS_S2_SR_HARMONIZED). In this study, we processed the Sentinel-2 MSI L2A dataset in 2019–2022 on the GEE platform to calculate LCC, and the calculated LCC is resampled to a specific resolution using the nearest neighbor method. The product of Global Land Cover with a Fine Classification System at 30 m21 was used to define the vegetation types. Based on the GLC_FCS30 land-cover product, vegetation worldwide was reclassified into five major types: broadleaf forest, needleleaf forest, cropland, grassland, and shrub. Empirical regression relationships between LCC and CSI were constructed for each type.Data processingFigure 1 illustrates the diagram to generate the high-resolution LCC product and the validation procedure.Fig. 1Diagram of generating the LCC product.Full size imageFirstly, the Sentinel-2 MSI L2A product was selected to calculate CSI as the following equation,$${CSI}=2.5times frac{{rho }_{{NIR}}-{rho }_{{RE}1}}{{rho }_{{NIR}}+{rho }_{{RE}1}}times frac{{rho }_{{Blue}}}{{rho }_{{RE}1}}$$
    (1)
    where ({rho }_{{Blue}}), ({rho }_{RE1}), ({rho }_{N{IR}}) represents the reflectance of blue, red-edge band 1, and NIR band of Sentinel-2. CSI is derived from the product of NDVIre ((frac{{rho }_{{NIR}}-{rho }_{{RE}1}}{{rho }_{{NIR}}+{rho }_{{RE}1}})) and the (frac{{rho }_{{Blue}}}{{rho }_{{RE}1}}) factor. While NDVIre increases with both LCC and LAI, the (frac{{rho }_{{Blue}}}{{rho }_{{RE}1}}) factor rises with LCC but declines with LAI. Their product thus enhances sensitivity to LCC while reducing LAI interference19.During the calculation, the 20m-resolution red-edge bands of Sentinel-2 were resampled to 10 m using the nearest neighbour method. Meanwhile, the global vegetation cover map derived from the GLC_FCS30D product21 was resampled to 10 m resolution using the nearest neighbour method and then reclassified into five types in Table 1.Table 1 Empirical regression equation between CSI and LCC in different vegetation types.Full size tableSecondly, using the type-specific regression equations (Table 1), LCC for different vegetation types was calculated on the GEE platform. The type-specific regression equations were acquired using radiative transfer models, and the key parameters were set according to Table S2 in the Supplementary Information. The RMSE of the regression model between CSI and LCC was between 6.04 and 10.21 μg/cm2 and the R2 is between 0.68 and 0.99 (Table 1). A cloud score based on the Sentinel-2 MSI was used to evaluate the cloud possibility, and the cloud-contaminated pixels can be identified using the algorithm (https://github.com/openforis/gee-gateway/blob/master/gee_gateway/gee/utils.py#L691). To make the max value of the product 80 μg/cm2, any value larger than 80 μg/cm2 was set to 80 μg/cm2. In this way, 1-day LCC maps with the same resolution as Sentinel-2 MSI (10 m) were generated.Thirdly, several 1-day LCC maps of the same tiles were averaged for different composite requirements from customers, and subsequently, resampled to produce the LCC product with customized spatial resolutions. The method to generate the high-resolution LCC product is available on the GEE platform (see the Supplementary information).Finally, the product’s retrieval rate (RR) in different day-composite strategies was calculated to evaluate the missing rates of the product. Additionally, the products in different resolutions were validated using ground-measured data and compared with the MODIS LCC product.The large volume of global high-resolution product data requires substantial storage space, and uploading such data can be extremely time-consuming. Therefore, we only uploaded the global 100 m/10 days resolution LCC product from 2019 to 2023 to the online server. For the higher-resolution LCC product, we provided the code and a web interface based on the GEE platform (https://code.earthengine.google.com/a06dfc261ad8019e025153d5bd0e68ca), allowing users to independently select their desired temporal and spatial ranges as well as the corresponding resolutions.Data RecordThe dataset (MuSyQ Global LCC product with 100 m/10 days resolution) is available at Science Data Bank22,23,24,25,26, and the link for each year’s product is shown in Table 2. Each year’s data is stored in a folder named after the corresponding year (e.g., ‘2019’). These year folders contain multiple subfolders organized by latitude and longitude (e.g., ‘E0N10’), within which the LCC product images are stored in *.tif format, named as ‘LCC_longitude-latitude_day-of-year.tif’ (e.g., ‘LCC_E0N10_001.tif’). The mean size of each file was 10.76 MB. The dataset is licensed under CC BY 4.0. For the higher-resolution product or product of customized spatial and temporal resolutions and ranges, users can download it using the GEE-based link (https://code.earthengine.google.com/a06dfc261ad8019e025153d5bd0e68ca). For detailed instructions, please refer to the Section 3 of the Supplementary Information.Table 2 Links of the MuSyQ Global Leaf Chlorophyll Content product with 100 m/10 days resolution.Full size tableTechnical ValidationValidation methodsThe ground-measured LCC from different research is collected to validate the MuSyQ Global LCC product. Details of the ground-measured data are shown in the Ground Measurement section of the Supplementary Information. These data encompassed 1139 sampling measurements in different field campaigns, including the National Ecological Observatory Network (NEON) in the USA, Huailai and Gaocheng field experiments in China. The smallest size of each experiment is 20 m * 20 m, which can validate the product with 10 m spatial resolution. To minimize temporal mismatch, we compared the 10-day mean composites derived from 10m-resolution MuSyQ Global LCC product with ground-based observations collected during the closest possible time windows. As for the validation of the 100m-resolution and the 500m-resolution LCC products, the 30m-resolution land cover product (GLC_FCS30) was used to assess if plots in Huailai and NEON are located in a homogeneous area. The centre coordinates of each plot in the experiment and its vegetation type in GLC_FCS30 were extracted. For the validation of the 100 m resolution product, sampling points are selected based on the presence of data with the same GLC_FCS30 product value within a 5 * 5 pixel area (i.e., 150 m * 150 m homogeneous). For the validation of the 500 m resolution product, sampling points are chosen within a 17*17 pixel area that shares the same GLC_FCS30 product value (i.e., 510 m * 510 m homogeneous).Direct validationFigure 2 illustrates the validation result of the 10m-resolution MuSyQ Global LCC product. The results suggest the overall RMSE and rRMSE of the product in 5 different vegetation types were 13.69 μg/cm2, 33.70%, respectively, and the R2 was 0.37. The LCC retrieved from the 10m-resolution product and the ground-measured LCC were aligned along the 1:1 line, with underestimation under high LCC conditions and overestimation under low LCC conditions. Figure 2b–f shows the detailed results in different vegetation types. The accuracy of the retrieved LCC is varied, with an RMSE between 11.28 and 15.22 μg/cm2 in the five types. The cropland had the highest accuracy with an RMSE of 11.28 μg/cm2, rRMSE of 19.36%, and bias of 1.57 μg/cm2. The grassland had the accuracy with an RMSE of 11.93 μg/cm2, rRMSE of 35.30%, and bias of 1.25 μg/cm2. The RMSE, rRMSE, and bias of the broadleaf forest were higher, with 14.25 μg/cm2, 36.24%, and 3.97 μg/cm2, respectively. LCC of needle leaf forest had the lowest accuracy, with RMSE of 15.22 μg/cm2. Figure 2b–f also shows LCC product for all five types tended to be underestimated when LCC is more than 60 μg/cm2 and the LCC was overestimated for forests, grasslands, and shrubs when LCC is less than 30 μg/cm2.Fig. 2(a) Validation of the 10m-resolution MuSyQ Global LCC; Validation results of the 10m-resolution MuSyQ Global LCC in broadleaf forest (b), needleleaf forest (c), cropland (d), grassland (e), and shrub (f).Full size imageFigure 3 shows the product’s bias under different LCC conditions. Due to the limited number of ground-measured LCC of shrubs, only the other four types were compared. Generally, the LCC retrieved from the product tended to be overestimated when the LCC was less than 20 μg/cm2. The bias of broadleaf forest, needleleaf forest, and grassland was more than 10 μg/cm2. When the LCC was 20 – 40 μg/cm2, the overestimation became less obvious, especially for the cropland, whose bias is close to 0 μg/cm2. When LCC increased, the overestimation gradually turned to underestimation for the broadleaf, needleleaf, and grassland. When the ground-measured LCC was 60 – 80 μg/cm2, the bias of the product for broadleaf forest and grassland declined to below 15 μg/cm2. As for the bias of cropland, it fluctuated with the increase in LCC. Apart from the condition of LCC = 40 – 60 μg/cm2, the mean bias of the product is close to 0 μg/cm2, indicating that the overestimation or underestimation is relatively slight.Fig. 3Bias of the LCC product under different LCC conditions. Due to the limited number of shrubs, only the first four vegetation types are compared. The black line within each violin represents the mean value of the bias, and the box represents the value of the upper and lower quartiles.Full size imageThe published 100m-resolution MuSyQ Global LCC product was validated using the 361 ground measurements (Fig. 4a and Table 3). The accuracy (RMSE = 15.11 μg/cm2, bias = 3.69 μg/cm2, R2 = 0.15) was lower than the 10m-resolution product for all five vegetation types. The grassland had the highest accuracy with the RMSE and bias of 11.19 μg/cm2, and −0.19 μg/cm2. For the 500m-resolution product, which shares the same resolution as the MODIS LCC product, its overall accuracy was lower than that of the 100m-resolution product, meaning that the resolution is an important factor contributing to the accuracy of the LCC product. RMSE rose to 15.77 μg/cm2 and the overall bias was 6.02 μg/cm2. The overestimation existed for the broadleaf forest, the needleleaf forest. Due to limited validation samples for cropland and shrub at the 500 m resolution, the current accuracy assessment for these vegetation types has relatively low representativeness. A more robust evaluation will require additional ground-based measurements.Fig. 4Validation of the 100m-resolution (a) and the 500m-resolution MuSyQ Global LCC product (b) in different vegetation types.Full size imageTable 3 Accuracy of 100m-resolution LCC product in different vegetation types.Full size tableSpatial patterns of MuSyQ Global LCCFigure 5a,b illustrate the global distribution of the MuSyQ LCC product in January and July 2020. The line chart (Fig. 5c,d) illustrate the variation in average LCC across different latitudes. Overall, global LCC was lower in January compared with July. In the region between 0–30°S, LCC was relatively high, with an average close to 40 μg/cm2. The highest LCC values were observed in the mid to low-latitude regions of eastern South America and Africa. Additionally, the low-latitude areas of the northern hemisphere, such as the Indian subcontinent, also exhibited high LCC values. In the other northern hemisphere regions, LCC was generally below 30 μg/cm2 due to the winter season. In July, the average LCC in the 30°N–60°N region was above 20 μg/cm2, and in the 45°N–75°N region, it was generally above 30 μg/cm2. The map showed that the highest LCC values were found in the northern parts of the Eurasian continent and the mid-latitude eastern regions of North America. In contrast, the Southern Hemisphere generally exhibited lower LCC values, mostly below 20 μg/cm2. Figure 5e illustrates the detailed information of typical vegetation sites.Fig. 5Global distribution of the MuSyQ Global LCC product. The product was resampled to 1 km, and the mean values of the LCC products in January 2020 (a) and July 2020 (b) were shown in the map. The line charts (c) and (d) represent the averaged LCC in different latitudes. (e) represented the details of specific sites for different vegetation types across the year 2020.Full size imageConsistency assessment between MODIS LCC and MuSyQ Global LCCFigure 6 illustrates the differences between the MuSyQ Global LCC and MODIS LCC. In January, the MuSyQ Global LCC product was generally lower than the MODIS product in the southern hemisphere, while in the mid to high-latitude regions of the northern hemisphere, the MuSyQ Global LCC was slightly higher than the MODIS LCC. The histogram on the right shows that the distribution of LCC’s differences (ΔLCC = MuSyQ Global LCC – MODIS LCC) was concentrated in the negative value region, indicating that the values of the MuSyQ Global LCC in January were lower than those of the MODIS LCC. Additionally, ΔLCC for most pixels was within ±5 μg/cm2, suggesting good consistency between the two products. Figure 6b shows the spatial distribution of the differences between the two products during July. The regions with the largest ΔLCC were in the northeastern part of the Eurasian continent, northeastern North America, southern Africa, and the eastern regions of the southern hemisphere. The regions with the smallest ΔLCC were in the western part of the Eurasian continent, the Northern Hemisphere regions of Africa, and western Australia. The histogram on the right indicates that in July, the peak of ΔLCC was around 0 μg/cm2, and the distribution of the histogram was more symmetrical compared to January, with no significant tendency towards the negative value region.Fig. 6Difference between MuSyQ Global LCC and MODIS LCC in January (a) and July (b) 2020.Full size imageDue to the temporal overlap of the MODIS LCC product (2000–2020) and available Sentinel-2 imagery (L2A level, after 2019) on the GEE platform being limited to 2019 and 2020, the study selected validation points from these two years for an intercomparison between the products (Fig. 7). There were 57 validation points, including where the two 500 m products overlapped from 2019 to 2020. The accuracy of the two products was similar, with the MuSyQ LCC product showing slightly higher accuracy than the MODIS LCC product. The RMSE and bias improved from 14.74 μg/cm2 and −2.65 μg/cm2 to 14.16 μg/cm2 and 1.68 μg/cm2, respectively. In the 500 m scale, the two products tended to show an obvious underestimation under high LCC conditions. The MuSyQ Global LCC product exhibited lower RMSE for broadleaf forests and grasslands. In comparison, the RMSE for needleleaf forests was higher than that of the MODIS product.Fig. 7Validation of the 500m-resolution MuSyQ Global LCC product and 500m-resolution MODIS LCC.Full size imageTemporal profiles of global LCC productsFigure 8 compares the temporal profiles of the four specific vegetation types. Both products effectively captured the phenological characteristics of typical vegetation types, showing an initial increase followed by a decrease in LCC. The MuSyQ Global LCC product generally fluctuated between adjacent time, while the MODIS LCC time series was smooth. Additionally, the 10m-resolution MuSyQ Global LCC showed higher values than the 500m-resolution MODIS LCC for all four types during the summer, which is the primary difference between the two products. For broadleaf forests, needleleaf forests, and crops, the MuSyQ Global LCC reached values above 70 μg/cm2 during the summer, while the maximum values of the MODIS LCC were all below 60 μg/cm2. For grasslands, the maximum value of the MuSyQ Global LCC approached 50 μg/cm2, whereas the MODIS LCC was below 40 μg/cm2. The MuSyQ LCC product exhibits stronger seasonal variability compared to MODIS LCC, likely because MODIS LCC undergoes temporal reconstruction, which smooths its time series. Additionally, non-vegetation components within mixed pixels can further reduce LCC values. Future work will explore time-series smoothing methods for high-resolution MuSyQ Global LCC.Fig. 8Temporal profiles of LCC in (a) broadleaf forest; (b) needleleaf forest; (c) cropland; (d) grassland.Full size imageRetrieval rate of MuSyQ global LCCFigure 9 presents the RR of the product under different temporal composition strategies. RR continuously increased with the extension of composition days. The 10-day composite product exhibited an RR of 33.2%–70.7%, the 20-day composite product reached 42.8%–79.9%, and the 30-day composite product further improved to 47.9%–83.4%. Additionally, in terms of RR across different seasons, winter and spring exhibited lower RR values, mostly below 60%, while summer showed higher RR values, with the 20-day and 30-day composite products generally exceeding 70%.Fig. 9Retrieval rate (RR) of the MuSyQ Global LCC under different temporal compositing strategies in 2020.Full size imageUncertainties of MuSyQ Global LCCHigh-resolution satellites typically have limited swath widths, resulting in longer revisit periods. The revisit period of Sentinel-2 is 5 days; issues such as cloud and rain cover can lead to high pixel missing in the 10-day composite product (Fig. 9). In the future, the reconstruction algorithm of the time series for the LCC will further enhance the applicability of the MuSyQ Global LCC product. Additionally, the MuSyQ Global LCC product utilizes the blue band, which is sensitive to atmospheric aerosols. Consequently, variations in atmospheric conditions during Sentinel-2 imaging can lead to fluctuations in LCC values, resulting in uneven brightness patterns. More accurate atmospheric correction and cloud masking algorithms will further enhance the accuracy of this product.

    Code availability

    The code to generate the high-resolution LCC products can be accessed using the following link: https://code.earthengine.google.com/a06dfc261ad8019e025153d5bd0e68ca. This code is written in JavaScript and can be executed directly on the GEE platform. Logging into an individual GEE account is necessary.
    ReferencesLu, X., Croft, H., Chen, J. M., Luo, Y. & Ju, W. Estimating photosynthetic capacity from optimized Rubisco–chlorophyll relationships among vegetation types and under global change. Environ. Res. Lett. 17, 014028 (2022).Article 
    ADS 
    CAS 

    Google Scholar 
    Gitelson, A. A., Gritz, Y. & Merzlyak, M. N. Relationships between leaf chlorophyll content and spectral reflectance and algorithms for non-destructive chlorophyll assessment in higher plant leaves. J. Plant Physiol. 160, 271–282 (2003).Article 
    CAS 
    PubMed 

    Google Scholar 
    Luo, X. et al. Incorporating leaf chlorophyll content into a two-leaf terrestrial biosphere model for estimating carbon and water fluxes at a forest site. Agric. For. Meteorol. 248, 156–168 (2018).Article 
    ADS 

    Google Scholar 
    Luo, X., Croft, H., Chen, J. M., He, L. & Keenan, T. F. Improved estimates of global terrestrial photosynthesis using information on leaf chlorophyll content. Glob. Change Biol. 25, 2499–2514 (2019).Article 
    ADS 

    Google Scholar 
    Croft, H. et al. The global distribution of leaf chlorophyll content. Remote Sensing of Environment (2020).Xu, M. et al. A 21-year time series of global leaf chlorophyll content maps from MODIS imagery. IEEE Transactions on Geoscience and Remote Sensing 60, 1–13 (2022).
    Google Scholar 
    Qian, X. et al. Global Leaf Chlorophyll Content Dataset (GLCC) from 2003–2012 to 2018–2020 Derived from MERIS and OLCI Satellite Data: Algorithm and Validation. Remote Sensing 15, 700 (2023).Article 
    ADS 

    Google Scholar 
    Xu, M. et al. Retrieving global leaf chlorophyll content from MERIS data using a neural network method. ISPRS Journal of Photogrammetry and Remote Sensing 192, 66–82 (2022).Article 
    ADS 

    Google Scholar 
    Wang, X. et al. Intercomparison and validation of five existing leaf chlorophyll content products over China. Int. J. Appl. Earth Obs. Geoinformation 130, 103930 (2024).Article 

    Google Scholar 
    Yu, W. et al. Global Land Cover Heterogeneity Characteristics at Moderate Resolution for Mixed Pixel Modeling and Inversion. Remote Sensing 10, 856 (2018).Article 
    ADS 

    Google Scholar 
    Yang, F., Ren, H., Li, X., Hu, M. & Yang, Y. Assessment of MODIS, MERIS, GEOV1 FPAR Products over Northern China with Ground Measured Data and by Analyzing Residential Effect in Mixed Pixel. Remote Sens. 6, 5428–5451 (2014).Article 
    ADS 

    Google Scholar 
    Li, J. et al. Can 3D model improve the accuracy of leaf chlorophyll content estimation using UAV and sentinel-2 data? Int. J. Appl. Earth Obs. Geoinformation 143, 104810 (2025).Chu, T. et al. Regional Analysis of Dominant Factors Influencing Leaf Chlorophyll Content in Complex Terrain Regions Using a Geographic Statistical Model. Remote Sens. 16, 479 (2024).Mirhoseini Nejad, S. M., Abbasi-Moghadam, D. & Sharifi, A. ConvLSTM–ViT: A Deep Neural Network for Crop Yield Prediction Using Earth Observations and Remotely Sensed Data. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 17, 17489–17502 (2024).Article 
    ADS 

    Google Scholar 
    Sharifi, A. & Safari, M. M. Enhancing the Spatial Resolution of Sentinel-2 Images Through Super-Resolution Using Transformer-Based Deep-Learning Models. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 18, 4805–4820 (2025).Article 
    ADS 

    Google Scholar 
    Schimel, D. et al. Observing terrestrial ecosystems and the carbon cycle from space. Glob. Change Biol. 21, 1762–1776 (2015).Article 
    ADS 

    Google Scholar 
    Guan, L. et al. A dataset of 30 m/10-day spatio-temporal continuous leaf chlorophyll content of MuSyQ GF-series (2021–2022, China). https://doi.org/10.57760/sciencedb.16673 (2025).Li, J. et al. MuSyQ 30m/10days leaf chlorophyll content product (from 2019 to 2020 across China version 01). https://doi.org/10.11922/sciencedb.j00001.00265 (2021).Zhang, H. et al. A novel red-edge spectral index for retrieving the leaf chlorophyll content. Methods in Ecology and Evolution 13, 2771–2787 (2022).Article 

    Google Scholar 
    Gu, C. et al. Retrieving decametric-resolution leaf chlorophyll content from GF-6 WFV by assessing the applicability of red-edge vegetation indices. Comput. Electron. Agric. 215, 108455 (2023).Zhang, X. et al. GLC_FCS30: global land-cover product with fine classification system at 30m using time-series Landsat imagery. Earth Syst. Sci. Data 13, 2753–2776 (2021).Article 
    ADS 

    Google Scholar 
    Zhang, H. et al. MuSyQ Global Leaf Chlorophyll Content product with 100m/10 days resolution (2019). https://doi.org/10.57760/sciencedb.19595 (2025).Zhang, H. et al. MuSyQ Global Leaf Chlorophyll Content product with 100m/10 days resolution (2020). https://doi.org/10.57760/sciencedb.19687 (2025).Zhang, H. et al. MuSyQ Global Leaf Chlorophyll Content product with 100m/10 days resolution (2021). https://doi.org/10.57760/sciencedb.19689 (2025).Zhang, H. et al. MuSyQ Global Leaf Chlorophyll Content product with 100m/10 days resolution (2022). https://doi.org/10.57760/sciencedb.19691 (2025).Zhang, H. et al. MuSyQ Global Leaf Chlorophyll Content product with 100m/10 days resolution (2023). https://doi.org/10.57760/sciencedb.19692 (2025).Download referencesAcknowledgementsThis work was supported by the National Key Research and Development Program (Grant No. 2023YFF1303602), and the Postdoctoral Fellowship Program of China Postdoctoral Science Foundation (CPSF) (Grant No. GZC20232759), and the National Natural Science Foundation of China (42271359).Author informationAuthors and AffiliationsState Key Laboratory of Remote Sensing and Digital Earth, Aerospace Information Research Institute, Chinese Academy of Sciences, Beijing, 100101, ChinaHu Zhang, Jing Li, Li Guan, Yadong Dong & Jing ZhaoUniversity of Chinese Academy of Sciences, Beijing, 100049, ChinaJing Li, Chenpeng Gu, Li Guan, Xiaohan Wang, Faisal Mumtaz & Qinhuo LiuNational Engineering Research Center for Satellite Remote Sensing Applications, Chinese Academy of Sciences, Beijing, 100101, ChinaChenpeng Gu, Xiaohan Wang, Faisal Mumtaz & Qinhuo LiuCarbon-Water Research Station in Karst Regions of Northern Guangdong, School of Geography and Planning, Sun Yat-Sen University, Guangzhou, 510006, ChinaShangrong LinCenter for Geospatial Information, Shenzhen Institute of Advanced Technology, Chinese Academy of Science, Shenzhen, 518055, ChinaWentao YuAuthorsHu ZhangView author publicationsSearch author on:PubMed Google ScholarJing LiView author publicationsSearch author on:PubMed Google ScholarChenpeng GuView author publicationsSearch author on:PubMed Google ScholarLi GuanView author publicationsSearch author on:PubMed Google ScholarXiaohan WangView author publicationsSearch author on:PubMed Google ScholarFaisal MumtazView author publicationsSearch author on:PubMed Google ScholarYadong DongView author publicationsSearch author on:PubMed Google ScholarJing ZhaoView author publicationsSearch author on:PubMed Google ScholarQinhuo LiuView author publicationsSearch author on:PubMed Google ScholarShangrong LinView author publicationsSearch author on:PubMed Google ScholarWentao YuView author publicationsSearch author on:PubMed Google ScholarContributionsH.Z. designed the algorithm and wrote the original manuscript. J.L. and Q.L. organized the generation of the product. C.G., X.W. generated the product, and L.G. developed the web interface. F.M. helped to revise the original manuscript. Y.D. and J.Z. organized the experiments, and S.L., W.Y. carried them out.Corresponding authorsCorrespondence to
    Jing Li or Qinhuo Liu.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationRights and permissions
    Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
    Reprints and permissionsAbout this articleCite this articleZhang, H., Li, J., Gu, C. et al. A high-resolution global leaf chlorophyll content product using the Sentinel-2 data.
    Sci Data 12, 1997 (2025). https://doi.org/10.1038/s41597-025-05831-xDownload citationReceived: 20 June 2025Accepted: 18 August 2025Published: 24 December 2025Version of record: 24 December 2025DOI: https://doi.org/10.1038/s41597-025-05831-xShare this articleAnyone you share the following link with will be able to read this content:Get shareable linkSorry, a shareable link is not currently available for this article.Copy shareable link to clipboard
    Provided by the Springer Nature SharedIt content-sharing initiative More