More stories

  • in

    Scale morphometry, geometry and ultrastructure of three Nemipterus species from the Egyptian part of the Red Sea

    AbstractThe present investigation aimed to document and analyze the diversity of scale characteristics in three Nemipterus species; N. zysron (12.2–18.5 cm SL), N. randalli (13–18 cm SL), and N. japonicus (8.5–17.2 cm SL) collected from the Egyptian Red Sea near Hurghada. Assessing interspecific differences in scale morphology, geometry, and morphometry provides valuable insights for taxonomy and stock identification. The results revealed pronounced interspecific variations in scale geometry, morphometrics, and radii meristics, particularly with respect to overall shape and size. Detailed structural features and surface ornamentation were examined using light microscopy and scanning electron microscopy. Notable differences were observed in surface morphology, interradial and intercircular grooves, interradial tongues and circuli, denticles, inner and outer lateral circuli and caudal field segmentation and granulation pattern. The observed variation in scale form among the three species underscores the potential utility of scale morphology in stock discrimination. Collectively, these findings contribute to improved species differentiation and offer a valuable tool for fisheries management and taxonomic assessment.

    Similar content being viewed by others

    Ruminant inner ear shape records 35 million years of neutral evolution

    Article
    Open access
    06 December 2022

    Self-organized patterning of crocodile head scales by compressive folding

    Article
    Open access
    11 December 2024

    A starfish-inspired 4D self-healing morphing structure

    Article
    Open access
    25 September 2024

    IntroductionFish scales, composed of ossified platelets, function as rigid external structures exhibiting complex morphological and structural characteristics. These features include scale type, shape, size, and specific surface ornamentations such as circuli, radii, lepidonts, and granules1. Because scales grow proportionally with the fish, new circuli are added at the periphery and remain preserved throughout the lifespan of the scale2,3. Scale shape is generally species-specific and has therefore been widely applied in elucidating fish systematics, determining stock structure, and assessing evolutionary relationships4,5. Considerable variability in scale morphology has been documented among different fish species, reflecting both intraspecific morphological plasticity and regional variation within individual specimens3,6. Such variation in scale shape may also contribute to reducing frictional drag during locomotion1.In addition to their protective role against external injuries, certain diseases and predation, scales can influence swimming performance. Surface features such as ctenii, circuli, radii, and scale curvature affect the flexibility and rigidity of the scales, thereby influencing the magnitude of force required for body bending during locomotion and ultimately shaping overall swimming efficiency2,5,7.Scale extraction is a non-lethal procedure that enables easy acquisition, preservation, and preparation for image-based analyses1. Owing to their distinctive morphological and structural characteristics, scales provide valuable information across a wide range of scientific applications, including fisheries assessment, age and growth determination, taxonomy, evolutionary and phylogenetic studies, archaeology and environmental monitoring1,2,3,5,8,9,10,11,12,13,14.Numerous studies have applied a variety of methods to examine the morphometrics, geometric properties, morphological characteristics, and surface ornamentations of fish scales in both marine and freshwater species1,3,4,6,7,8,11,15,16,17,18,19,20.The family Nemipteridae (threadfin breams) comprises approximately 77 species across five genera21 and is characterized by small to moderate body size and vibrant coloration22. Members of this family are primarily distributed throughout the tropical and subtropical Indo-West Pacific but are absent from the eastern Pacific and Atlantic Oceans23. The largest genus, Nemipterus23, includes many species of economic importance as a food source24. Extensive research on Nemipterus has addressed its biology24,25,26,27,28,29, population dynamics and fisheries30,31,32,33 as well as its biochemistry34 species identification and genetic diversity23,35,36,37.Since the opening of the Suez Canal, the Mediterranean Sea has become connected to the Red Sea, facilitating a substantial influx of tropical fauna into the predominantly eastern Mediterranean basin. Numerous Indo-Pacific species (Lessepsian migrants) have consequently established populations and expanded their ranges within the region38. Nemipterus japonicus and N. randalli have been documented in the Mediterranean Sea39,40.Threadfin breams constitute an important component of coastal demersal fish communities and regional fisheries41,42. In recent years, they have emerged as a significant fishery resource in Egyptian waters, particularly in the Suez Gulf trawl fishery, where their catch now accounts for approximately 7% of the total trawl yield43.Despite the economic importance of Nemipterus species in Egypt, their scale characteristics have been relatively understudied. The present study aimed to investigate scale traits that can differentiate three nemipterid species; Nemipterus zysron, N. randalli, and N. japonicus from the Red Sea near Hurghada. Scales were examined using light and scanning electron microscopy, and univariate and multivariate analyses were applied to assess morphometric and geometric variations in size and shape.Materials and methodsSpecimen collectionA total of 601 scales were collected from three Nemipterus species: N. zysron (12.2–18.5 cm standard length), N. randalli (13–18 cm standard length), and N. japonicus (8.5–17.2 cm standard length). Specimens were obtained from fish markets in Hurghada (27° 15′ N, 33° 48′ E; Red Sea, Egypt) between September 2021 and August 2022.Preparation and measuring of scalesScales were carefully extracted from the left side of the body, targeting seven specific anatomical regions as shown in Fig. 1a: (A) beneath the anterior portion of the dorsal fin (BDFS), (B) post-operculum (POS), (C) beneath the lateral line, between the pectoral and pelvic fins (BLLS), (D) caudal peduncle (CPS), (E) anterior lateral line (ALLS), (F) middle lateral line (MLLS), and (G) posterior lateral line (PLLS). Extracted scales were thoroughly cleaned in a 10% ammonia solution for 24–36 h to remove adhering tissue without compromising surface microstructure. Following cleaning, scales were dried on filter paper and mounted between two glass slides for storage.Scales obtained from the first four regions were used to count primary (R1) and secondary (R2) radii and to obtain morphometric measurements, including scale width (W), caudal field length (L2), rostral field length (L1), and scale length (L) (Fig. 1b), for assessing stock-related differences among species. The following morphometric indices were subsequently calculated and evaluated: R1/W, W/L, W/L1, W/L2, L1/L, L2/L, and L1/L2. To assess variation associated with the lateral line canal, scales from the remaining regions were analyzed.Scale landmark coordinationsIn order to distinguish between distinct fish stocks, scales collected from the region beneath the pectoral fin of the Nemipterus species under study were examined using geomorphometric techniques. Eight homologous landmarks were digitized on consistent anatomical positions of each scale using tpsUtil and tpsDig2 software44. These landmarks included the ventro- and dorso-lateral tips of the rostral margin (landmarks 1 and 2), the dorso- and ventro-rostral centers (landmarks 3 and 7), the dorsal and ventral tips of separation line (landmarks 4 and 6), the posterior tip of the scale (landmark 5) and the focal point of the scale (landmark 8) (Fig. 1b).Fig. 1(a) Photograph of N. randalli, a representative of the studied species, illustrating the body regions from which scales were collected (A: BDFS, B: POS, C: BLLS, D: CPS, E: ALLS, F: MLLS, G: PLLS). (b) Schematic representation of a scale illustrating the primary radii (R1), secondary radii (R2), morphometric measurements (W: scale width, L: scale length, L1: rostral field length and L2: caudal field length) and landmark positions.Full size imageMicroscopic studyScales from the first four body regions were photographed using a stereomicroscope equipped with an AxioCam ERc 5s camera (Carl Zeiss–Promenade 10; 07745 Jena, Germany) and operated with Zeiss imaging software at the National Institute of Oceanography and Fisheries. Scales from all regions were subsequently examined using scanning electron microscopy (SEM) to assess their microstructural characteristics. For SEM preparation, cleaned and dried scales were mounted on specimen stubs using adhesive tape and coated with a 30-nm layer of gold. Electron micrographs were obtained using a JEOL JSM-5400LV SEM operated in backscattered electron mode at an accelerating voltage of 15 kV at the Assiut University Electron Microscope Center, Assiut, Egypt. Interpretation of SEM features followed the descriptions and criteria outlined in Hassanien45, Sadeghi et al.13, Al Jufaili et al.1, and Mekkawy, et al.3, the SEM results were interpreted and described.Statistical analysisIBM SPSS version 2646 was used to compute descriptive statistics of the scale morphometric indices. A two-way analysis of variance (ANOVA; design: species + scale location + species × scale location) was performed to evaluate the effects of species and scale location. Post hoc comparisons were conducted using Tukey’s HSD test to assess interspecific variation among Nemipterus species. A multivariate analysis of variance (MANOVA) was also applied. In terms of homogeneity of variances, Levene’s test revealed significant findings for every variable.A two-way permutational multivariate analysis of variance (PERMANOVA) was conducted using the PAST version 4.11 software47 to further examine interspecific variation based on multivariate distances (Euclidean Index, 9,999 permutations). Furthermore, principal component analysis (PCA), discriminant function analysis (DFA), and cluster analysis based on Mahalanobis distances were performed Classification accuracy in DFA was assessed using leave-one-out cross-validation. Frequency distributions of radii counts across species and sampling locations were compared using Chi-square tests (Pearson Chi-square and likelihood ratio) to detect significant intra- and interspecific differences among Nemipterus stocks.Geometric morphometric analyses were conducted using MORPHOJ version 1.07c48. Generalized Procrustes Analysis (GPA) was used to standardize landmark configurations by removing variation due to size, position, and orientation, generating a set of shape variables49. Multivariate analyses, including PCA and Canonical Variate Analysis (CVA), were performed to examine interspecific differences in shape. Deformed outline drawings were generated to depict the shape variations identified by the CVA, using the mean shape as a reference framework for comparative analysis. The Mahalanobis distance (P < 0.0001) were used to evaluate the significance of mean shape differences among species. To assess the effects of size and shape variation in fish scales, distinct multivariate regression analyses of Procrustes coordinates on centroid size were conducted following Cavalcanti, et al.50.The landmark coordinates for the three Nemipterus species were transformed into distance measurements using PAST software. Stock discrimination was further investigated using 28 landmark-transformed distances analyzed through DFA and Multivariate Analysis of Covariance (MANCOVA). These analyses were conducted using raw data, distance indices standardized by standard length (SL), and measurements corrected for Burnaby’s allometry and isometry, employing both PAST and SPSS packages.ResultsOverall structural characteristics of scalesThe scales of the studied species are predominantly ctenoid with a sectioned structure, characterized by well-defined radii. Simple scales, characterized by the absence or faint development of radii, were not observed. The scales exhibited variation in shape they were predominantly polygonal (pentagonal or hexagonal) or rectangular in form. Among the examined body regions, the largest scales were observed in the post-opercular (POS) region. Scales collected from the first three body regions were generally broad along the dorsoventral axis (Fig. 2a), whereas those from the caudal peduncle (CPS) were broader along the anteroposterior axis (Fig. 2b). Each scale comprised four distinct fields: dorsal, ventral, rostral (anterior), and caudal (posterior).Surface ornamentation across the scale fields was primarily characterized by circular patterns created by alternating ridges (circuli) intersected by distinct, deep, and narrow grooves known as radii, that extended radially from the focus toward the anterior margin. Depending on their position and point of termination, the radii were classified into two categories: primary and secondary (Fig. 1b). This configuration was clearly visible in most fields, except in the caudal field, where circuli were absent and replaced by granulated structures (Fig. 2). The scale focus was situated posteriorly. The rostral margin of the scales exhibited a wavy and striated appearance, contrasting with the crescent-shaped contour of the caudal margin. The dorsal and ventral fields consistently displayed a convex form across all examined body regions. Regenerated scales were also observed in various body regions, distinguished by the lack of surface ornamentation, especially in the central portion of the scale (Fig. 2c).Fig. 2Types of scales observed in N. japonicus, representative for the other Nemipterus species studied. (a) scales broad at dorsoventral axis, (b) scales broad at anteroposterior axis (c) regenerated scales. Arrows indicate granulation segments.Full size imageTraditional morphometrics and meristics of scaleThe number of R1 varied across different body regions, ranging from 4 to 10 in N. zysron and from 4 to 9 in both N. randalli and N. japonicus. In contrast, R2 counts ranged from 0 to 3 in N. zysron, 0 to 4 in N. randalli, and 0 to 1 in N. japonicus (Tables 1 and 2). A statistically significant correlation was observed between the number of R1 and body size in N. zysron (P < 0.01), suggesting that R1 development in this species is size-dependent. However, no significant correlation was observed between R1 count and body size in N. randalli and N. japonicus, indicating that this trait is not influenced by body size in these species. Conversely, a significant correlation between R2 count and fish size was found in N. japonicus, whereas no such relationship was detected in N. zysron or N. randalli (P > 0.01), implying that R2 count is independent of body size in the latter two species.Descriptive statistics (mean ± SD) and occurrence percentages of R1 and R2 counts across the first four body regions of the examined Nemipterus species are summarized in Tables 1 and 2. The distribution of R1 differed significantly among body regions in all species (P < 0.001) and exhibited interspecific variation within each region (P < 0.035). The distribution of R2 varied significantly among body regions in N. zysron and N. randalli (P < 0.001), but not in N. japonicus (P > 0.206). Interspecific differences in R2 distribution were also detected across all examined regions (P < 0.033).Table 1 Basic statistics (Mean ± SD) and percentage of occurrence of primary radii (R1) counts of the scales of the studied Nemipterus species.Full size tableTable 2 Basic statistics (Mean ± SD) and percentage of occurrence of secondary radii (R2) counts of the scales of the studied Nemipterus species.Full size tableThe basic statistical values of scale morphometric indices obtained from the first four body regions are shown in Table 3. ANOVA revealed that species, body region, and their interaction significantly influenced all scale morphometric indices (P < 0.001). The parameters W/L2, L1/L and L1/L2, in region BDFS, L2/L in region POS, L1/L, L1/L2, W/L, W/L1, and W/L2 in region BLLS, as well as W/L, W/L1, and W/L2 in region CPS exhibited the highest values for N. zysron. The greatest values for N. randalli were observed in region BDFS for W/L, W/L1 and L2/L; in region POS for W/L, W/L1, W/L2, L1/L and L1/L2; and in regions BLLS and CPS for L2/L. The highest values for N. japonicus were observed in W/L in region BDFS, L1/L and L1/L2 in region CPS, and R1/W across all regions.All body regions under examination showed interspecific differences in scale indices, with the exception of the BDFS region with regard to the W/L2 index. Moreover, intraspecific variations in scale indices were observed in all species examined. The PERMANOVA results on the traditional scale indices across different body regions of the three Nemipterus species indicated no significant effect of species (P > 0.01), whereas scale location and the interaction between species and scale location exhibited highly significant effects (P < 0.0001).Table 3 Descriptive statistics, mean ± standard deviation (SD), range and number of scales (N) of the scale’s morphometric indices from the first four body regions of the studied Nemipterus species.Full size tableFigure 3 illustrates the clustering of scale locations derived from the significant Mahalanobis distance matrix (P < 0.0001) of the examined scale indices. No distinct clustering pattern was observed in relation to species or body regions, as there was considerable overlap between the body regions of different species. Interspecific differences were evident, irrespective of scale location, as demonstrated by the Mahalanobis distances (P < 0.0001). N. japonicus formed a distinct cluster, whereas the remaining species showed no clear separation on the CVI (82.49%), with 50.9% of cross-validated grouped cases accurately classified.Fig. 3Clustering of the studied Nemipterus species examined, utilizing squared Mahalanobis distances. The species and region abbreviations are displayed alongside the scale morphometric indices on the Y-axis. (N.z: N. zysron, N.r: N. randalli, N.j: N. japonicus)Full size imageGeometric morphometric analysisPrincipal components analysis (PCA) of the Procrustes coordinates of scales located beneth the pectoral fin revealed interspecific variations among the Nemipterus species studied (Fig. 4a). Notably, N. zysron was positioned within the range of the other two overlapping species on PCI (29.5%). Canonical variate analysis (CVA) of these coordinates further demonstrated clearer separation of species with minimal interspecific overlap on CVI (65.08%) (Fig. 4b). The predicted group membership accuracy for the original cases of N. zysron, N. randalli, and N. japonicus was 96.7%, 93.3%, and 90.0%, respectively, with an overall mean accuracy of 93.3% for correctly classified original cases. In contrast, the average cross-validation accuracy for grouped cases was 80.0%, with correct classification rates of 76.7% for N. zysron, 83.3% for N. randalli, and 80.0% for N. japonicus. Furthermore, a statistically significant difference was observed among species based on the combined Procrustes coordinates after controlling for standard length (SL) (Wilks’ λ = 0.172, p < 0.0001, partial η² = 0.586).The DFA of the regression residuals of the Procrustes coordinates revealed an interspecific pattern of variation consistent with the previous analysis (Fig. 4c). The predicted group membership accuracy for the original cases of N. zysron, N. randalli, and N. japonicus was 96.7%, 93.3%, and 90.0%, respectively, resulting in an average accuracy of 93.3% for correctly classified original cases. In contrast, the average accuracy for cross-validated grouped cases was also 93.3%, with correct classification rates of 96.7% for N. zysron, 93.3% for N. randalli, and 90.0% for N. japonicus. Furthermore, a statistically significant difference was found among species based on the combined regression residuals after controlling for centroid size (Wilks’ λ = 0.174, p < 0.0001, partial η² = 0.583).The DFA of the principal components of regression residuals derived from the Procrustes coordinates exhibited an interspecific variation pattern that substantially coincided with the aforementioned patterns (Fig. 4d). The predicted group membership for the original cases of N. zysron, N. randalli, and N. japonicus is 96.7%, 93.3%, and 90.0%, respectively, resulting in a mean classification accuracy of 93.3% for the original grouped cases. An average of 81.1% of cross-validated grouped cases were correctly classified for N. zysron (80.0%), N. randalli (80.0%), and N. japonicus (83.3%). A statistically significant difference among species was observed in the combined principal components of regression residuals after adjusting for centroid size (Wilks’ A = 0.166, p < 0.0001, partial η2 = 0.592).The DFA of landmark-transformed distances revealed interspecific variation among the studied Nemipterus species. Raw landmark distances (without normalization) showed species separation with some overlap (CVI = 64.84%) (Fig. 5a). When indices were scaled relative to standard length (SL), species clusters separate more distinctly, especially between N. japonicus and the other two species (CVI = 67.81%) (Fig. 5b). After removing allometric effects, the species clusters remained partially overlapping; however, N. japonicus tended to form a distinct group. In contrast, controlling for isometric size scaling resulted in a clearer separation of N. japonicus, while N. randalli and N. zysron remained partially overlapping (Fig. 5c, d). Removing allometric effects further enhanced separation among species (Fig. 5c). In contrast, controlling for isometric size scaling (i.e., analyzing shape variation independent of overall size) resulted in minor overlap between N. randalli and N. zysron, while N. japonicus remained distinct from N. zysron (Fig. 5d). Nenipterus species exhibited average correct classification rates of 91.1%, 91.1%, 90%, and 92.2% for raw distance, indices, isometry, and allometry transformed distances, respectively. Conversely, the species exhibited means of 76.7%, 73.3%, 75.6%, and 76.7% for correctly classified cross-validated grouped cases based on raw distance, indices, isometry, and allometry transformed distances, respectively. Clustering utilizing Mahalanobis distance revealed the subsequent relationship patterns among Nemipterus species: N. zysron + N. randalli vs. N. japonicus for raw distance, distance indices and isometry and allometry transformed distances.In conclusion, significant interspecific variations in size, shape, and shape-size among the studied Nemipterus species were observed through both univariate and multivariate analyses of their traditional and landmark-based scale morphometric traits. Isometry and allometry transformed distances, revealed unique patterns of shape variation.Fig. 4(a) Distribution of scores for the first two principal components (PC I, PC II) derived from the Procrustes coordinates of scales across the species under study. (b) Scatter plot from Canonical Variable Analysis (CVA) showing the scale shape based on coordinates (CVI, CVII). (c) CVA scatter plot illustrates the regression residuals of the Procrustes coordinates. (d) Discriminant function analysis of the CVA scatter plot based on the principal components of regression residuals derived from the Procrustes coordinates. Nz: N. zysron, Nr: N. randalli, Nj: N. japonicus.Full size imageFig. 5Scatter plot of Canonical Variable Analysis (CVA) depicting the landmarks-transformed distances of the species studied from Hurghada, Red Sea, Egypt, based on scale morphology. Nz: N. zysron, Nr: N. randalli, and Nj: N. japonicus.Full size imageScanning electron microscope studies (SEM)Rostral fieldThe inter-radial tongues and the first circulusWithin the inter-radial space, the rostral margins of the scales exhibit tongue-like projections that lack circuli near the rim (Fig. 6). Two distinct forms of these projections, along with the first inter-radial circulus, are observed in the scales of the examined nemipterid species. The first form is characterized by conical tongue-like projections and a crescent-shaped first inter-radial circulus, which was identified in N. zysron (BDFS, POS, CPS, MLLS, and PLLS), N. randalli (BDFS), and N. japonicus (BDFS) (Fig. 6a). The second form exhibits conical tongue-like projections with a straight first inter-radial circulus (Fig. 6b), this form was recorded in N. zysron (BLLS and ALLS), N. randalli and N. japonicus (all regions except BDFS).Fig. 6SEM of the rostral field of N. japonicus scales, representing the other two studied species, illustrating the radii (R), inter-radial tongues (IRT), inter-radial circuli (IRC) and the first inter-radial circulus (1st IRC). (a) Conical tongues with a crescent-shaped first inter-radial circulus. (b) Conical tongues with a straight first inter-radial circulus.Full size imageRadial groovesThe radial grooves of the scales of the nemipterid species under study are categorized into two distinct forms (Fig. 7): Form 1: wide and deep grooves with a thin membrane-like structure (Fig. 7a). This form was recorded in N. zysron (BDFS, BLLS & PLLS), N. randalli (BDFS, BLLS, ALLS, MLLS & PLLS) and N. japonicus (all regions). Form 2: narrow and deep grooves with a thin membrane-like structure (Fig. 7b). This form was recorded in N. zysron (POS, CPS, ALLS & MLLS) and N. randalli (POS & CPS). In both types, the ridges extend almost to the edge of the groove.Fig. 7SEM of the radial grooves in the Nemipterus species under study. (a) Wide, deep grooves with a thin, membrane-like structure in N. randalli scales, representing the other two studied species. (b) Narrow, deep grooves with a thin, membrane-like structure in N. randalli scales, representing N. zysron.Full size imageThe inter-radial circuli, grooves and denticlesThe inter-circular grooves in the rostral field were narrow and deep. The circuli exhibited small denticles within the inter-radial space. Five distinct forms of denticles were recorded on the dorsal surfaces of the inter-radial circuli (Fig. 8): Form 1: The denticles are conical in shape, featuring recurved, hook-like teeth (Fig. 8a) recorded in N. zysron (BDFS, BLLS & CPS). Form 2: blunt denticles are (Fig. 8b) recorded in N. zysron (POS) and N. randalli (BDFS). Form 3: conical denticles are with unicuspid ends (Fig. 8c) recorded in N. zysron (ALLS, MLLS & PLLS), N. randalli (POS, ALLS, MLLS & PLLS) and N. japonicus (POS, CPS ALLS, MLLS & PLLS). Form 4: the denticles are irregular shape (Fig. 8d) recorded in N. randalli (CPS & BLLS). Form 5: the denticles are conical with bicuspid and unicuspid ends (Fig. 8e) recorded in N. japonicus (BDFS & BLLS).Fig. 8SEM of the inter-radial circuli, grooves and denticles of the three studied Nemipterus species; (a) conical denticles, featuring recurved, hook-like teeth in N. zysron (b) blunt denticles on N. randalli circuli as representative of N. zysron, (c) conical denticles with unicuspid end on N. japonicus circuli representing the other two species showing, (d) irregular denticles in N. randalli, (e) conical denticles with unicuspid and bicuspid ends in N. japonicus.Full size imageOuter lateral circuli, grooves and denticlesIn the studied Nemipterus species, the outermost lateral circuli display thin, flat, and wide grooves. These circuli either bear denticles (Fig. 9a), as observed in N. zysron (BDFS, POS, CPS, MLLS & PLLS) and N. randalli (BDFS, POS & BLLS), or are free of denticles (Fig. 9b) as documented in N. zysron (BLLS & ALLS), N. randalli (CPS, ALLS, MLLS & PLLS), and N. japonicus (all regions).Fig. 9SEM of the outer lateral circuli, grooves and denticles of the studied Nemipterus species; (a) The outer circuli are thin and bear denticles with flat and wide grooves in N. zysron as representative of N. randalli; (b) The outer circuli are thin and free of denticles with flat and wide grooves in N. japonicus representing the other two species.Full size imageInner lateral circuli, grooves and denticlesIn the three studied Nemipterus species, the inner lateral circuli display two types: either thin, bearing numerous denticles with flat, shallow, and wide grooves (Fig. 10a), observed in N. zysron and N. randalli (all regions) and N. japonicus (BDFS, BLLS, CPS, ALLS, MLLS & PLLS), or thick, bearing few denticles with flat, deeper, and narrower grooves (Fig. 10b), documented in N. japonicus (POS).Fig. 10SEM of the inner lateral circuli, grooves and denticles of the studied Nemipterus species; (a) the inner circuli are thin, bear many denticles with flat, shallow and wide grooves in N. zysron representing the other two species; (b) The inner circuli are thick, bear few denticles with flat deeper and narrower grooves in N. japonicus.Full size imageThe focus regionThe focus region in the three Nemipterus species appears in two forms: it is either oval-shaped and surrounded by oval ridges (circuli), as observed in N. randalli (BDFS, POS & BLLS) and N. japonicus (POS & BLLS) (Fig. 11a) or round-shaped and surrounded by round ridges, as documented in N. zysron (all regions), N. randalli (CPS) and N. japonicus (BDFS & CPS) (Fig. 11b).Fig. 11SEM of the focus region in the three studied Nemipterus species; (a) oval shaped focus in N. japonicus as representative of N. randalli, (b) round shaped focus in N. zysron representing the other two species.Full size imageCaudal fieldIn all the studied species, the caudal field area is weakly delineated from the anterior rostral field by the separation line, and the posterior rim exhibits a curved shape, featuring ctenii (Fig. 12a). The granulation area in the caudal field is devoid of circuli; yet, it features fine segments and is permeated by many pores throughout all regions (Fig. 12b).Based on the separation line, the shape of the caudal region and the posterior rim five distinct scale forms were identified in the three Nemipterus species studied (Fig. 13): Form 1: The separation line is almost straight and the posterior rim is irregular crescent resulting in an oval caudal area. Form 2: The separation line is convex and the posterior rim is V-shaped resulting in a nearly conical caudal area. Form 3: The separation line and the posterior rim are irregular resulting in an elongated irregular oval caudal area. Form 4: The separation line is irregular crescent and the posterior rim is rounded resulting in almost a parallelogram caudal area. Form 5: The separation line is almost straight and the posterior rim is rounded resulting in an oval caudal area.Fig. 12SEM of the caudal region of N. japonicus scales, representing the other two species. The caudal region has fine segment with pores (P), ctenii (C), focus (F), posterior margin (PM), rostral field (RF), caudal field (CF), lateral field (LF).Full size imageFig. 13Schematic diagrams illustrating the forms of scales identified in the studied Nemipterus species according to the separation line, the posterior rim and the shape of granulation area in the caudal field.Full size imageLateral line scalesIn the studied species, the lateral line scales generally feature a wide tube (canal) that runs parallel to the anteroposterior axis of the scale in a straight manner, while some other tubes are oriented obliquely to this axis (Fig. 14a). This canal exhibits a posterior opening positioned medially within the focus region on the inner surface of the scale and located at a considerable distance from the posterior margin. The anterior opening is similarly situated away from the anterior margin. Additionally, lateral line canal pores are present on both sides of the canal (Fig. 14b).Fig. 14SEM of N. japonicus scales, representing the other two species showing the lateral line canal characteristics including anterior opening (AO) and pores (P).Full size imageDiscussionScale morphometrics, geomorphometrics, and microstructural analyses have been widely utilized in comparative, phylogenetic, and stock identification studies of fish species2,3,13,15,18,20,25,51,52. The methodologies utilized by these authors often incorporate univariate and multivariate analyses3,5,8,20, scanning electron microscopy (SEM)1,3,8,13,18 and geomorphological analyses2,16,17,20.To date, only a limited number of studies have investigated the scale morphology of nemipterid species, collectively demonstrating that scale characteristics can serve as reliable indicators for distinguishing among Nemipterus stocks. Renjith, et al.51 analyzed morphological variation in the tenth lateral line scale of N. japonicus, N. bipunctatus and N. randalli using elliptical Fourier analysis, revealing distinct interspecific differences in scale shape and highlighting the taxonomic value of scale morphology within the genus. Devi, et al.7 further described the scales of N. japonicus as exhibiting clear morphological variations across three different body regions, while Ujjania and Jaiswar53 compared variations in scale size between N. japonicus, N. bipunctatus and N. randalli, reinforcing the diagnostic potential of scale-based traits.The predominant scale type among marine fishes is the elasmoid type, encompassing both cycloid and ctenoid scales7. In the present study, the scales of Nemipterus species were identified as ctenoid and sectioned in type, exhibiting distinct surface ornamentation. These observations corroborate the findings of Devi, et al.7, who characterized the scale morphology of several species including N. japonicus, and those of Lelli, et al.41, who examined N. randalli from the eastern Mediterranean Sea. In addition, our results are consistent with the general morphological patterns reported in various marine taxa1,3,13,18.According to Spinner, et al.54, the evolution of ctenoid scales was likely driven by mechanical interactions with the surrounding environment, offering an adaptive advantage by reducing physical damage. This adaptation is particularly relevant for demersal species such as Nemipterus, which inhabit areas where contact with the sea bottom, rocks, and corals is frequent. Thus, the presence of ctenoid scales in Nemipterus may represent an evolutionary response to the mechanical constraints of a demersal lifestyle. Moreover, beyond their protective function, ctenoid scales also contribute to hydrodynamic efficiency. Complementary to this, Harabawy, et al.11 reported that the epidermal surface of fishes, especially those inhabiting coral reefs, is exposed to frictional forces generated by water movement. By regulating turbulence within the boundary layer, ctenoid scales help maintain smooth water flow along the body surface, thereby reducing drag and enhancing swimming performance2. Furthermore, some fish species exhibit both cycloid and ctenoid scales, reflecting variations in scale morphology and functional adaptations9,13,15.Scale shape is also thought to influence swimming performance. For instance, Ibañez, et al.55 suggested that scales elongated along the anteroposterior axis may reduce thrust and water pressure generated during swimming. In contrast, scales that are comparatively wide along the dorsoventral axis and shorter in the anteroposterior axis are thought to be advantageous for a subcarangiform swimming mode. In the present study, scales of the latter form were predominantly observed in the first three body regions. However, due to the lack of detailed information regarding the swimming behavior and performance of nemipterid species, a potential association between scale shape and swimming mode cannot be formally evaluated. The caudal peduncle scales are clearly distinguishable from those of the other body regions with respect to their size (as indicated by the W/L ratio) and overall shape. The elongation of these scales along the anteroposterior axis suggests a functional role in enhancing swimming efficiency by modulating thrust and water pressure55. These findings are consistent with observations reported for cichlid species from Lake Tanganyika2.The current study, along with various other investigations3,12,45,56 supports the validity of scale radii-meristics distribution and morphometrics in the identification of fish species and stocks. The number and morphology of radii serve as valuable taxonomic characters for species identification1. In the present study, both primary and secondary radii exhibited interspecific differences in the studied species. The primary radii varied intraspecifically among body regions in all species and was size-dependent only in N. zysron. Whereas, the secondary radii showed intraspecific variations in N. zysron and N. randalli, with size-dependence observed solely in N. japonicus. Few studies have investigated such variations1,3,12,56,57. In addition, primary radii in Nemipterus species exhibited higher counts than secondary radii, a situation also recorded previously by different authors3,45,58. According to Raffealla and Nath59 the formation of primary, secondary, and tertiary radii is generally considered a growth-related phenomenon.Moreover, it was suggested that the radii exhibited a weak influence from genetic factors, suggesting that other environmental or physiological factors may play a more significant role. In this regard, the enhanced nutritional conditions of the fish may correlate with an increased number of radii60. Given that radii represent gaps in scale ossification that contribute to scale flexibility61 while maintaining protective functionality62, their variation could reflect adaptive responses to environmental conditions rather than strictly genetic determinants. However, due to the absence of data on body flexibility of Nemipterus, it is currently impossible to comment on such a relationship. In addition, Esmaeili and Gholami61 reported that the number of radii depends on the scale location on fish body. Our results revealed that the BDFS, POS, and CPS regions showed the greatest number of radii counts among the examined species. According to Wainwright and Lauder62, such elevated radii counts typically occur in body areas that are inherently curved, such as the dorsal region, or in those subjected to greater lateral bending during locomotion, such as the caudal peduncle and tail.In the present study, both intraspecific and interspecific relationships were reflected through quantitative, size-independent morphometric indices of the scales. The obtained results are consistent with those of Ujjania and Jaiswar53, who investigated N. japonicus and N. randalli. Furthermore, several studies have employed comparable morphometric indices of fish scales in other teleost species1,3,5,8,11,63,64. The observed morphological differentiation among the studied species may be attributed to environmental, geographical, and biological variations influencing phenotypic expression53. In contrast to the present findings, previous authors did not address the variability of primary radii in relation to scale width across different species and sampling locations.Numerous investigations have focused on the analysis of fish populations and species identification through the examination of scale geometry utilizing different scale landmarks in multivariate size and shape assessments2,3,4,17,19,20. Also, the geometric analysis is harmless, rapid, and cost-effective than genetic analysis16.Our results revealed clear species-specific scale shape differences. This was supported by high classification accuracies and statistical significance, demonstrating that the methods effectively discriminate species by scale shape variation. In addition, landmarks transformed distances, substantially improved species discrimination. N. japonicus consistently clusters apart, whereas N. zysron and N. randalli, though more similar, are better separated once size-related shape variation is accounted for. This highlights the essential role of morphometric data transformations for accurate species classification, especially in closely related taxa with subtle morphological differences. These results align with recent studies employing geometric morphometrics for species discrimination in other fish taxa2,17,20,65.​.Various studies have investigated the ultrastructure and surface ornamentation of scales in the rostral and caudal regions to distinguish between species1,3,8,13,15. In the present study, the initial inter-radial circulus exhibited either a crescentic or straight configuration, a feature that may represent species-specific variation. Similar observations have been documented in various fish taxa3,11,45,56. Correspondingly, Lippitsch66 noted that the morphology of the first inter-radial circuli is characteristic within cichlid species, although it may occasionally be influenced by environmental conditions. Furthermore, the shape of the first circuli has been recognized as a potentially valuable taxonomic character in certain groups, such as goatfishes (Mullidae)15.Previous studies have reported that denticles exhibit considerable variation in shape and size across different fish species1,3,8,15,18. Variations in denticle morphology have been shown to depend on species identity, body size, and the specific location of the scales on the body6,15,18. The present investigation did not reveal significant interspecific differences in denticle morphology among the three Nemipterus species examined. This finding aligns with the observations of Al Jufaili, et al.1, who indicated that the taxonomic significance of denticles in Garra shamal remains uncertain, and with Gholami, et al.67, who reported that the microstructure of scales, particularly with respect to denticles, is insufficient for distinguishing among aphaniid species but may assist in differentiating certain populations. Conversely, Ferrito, et al.68 demonstrated that denticle morphology remains constant throughout the life of the fish and represents a reliable taxonomic feature for species identification. This is further supported by Lippitsch66 who reported that denticle morphology may serve as a useful character in phylogenetic analyses.Moreover, the presence of denticles on the inter-radial circuli and rostrolateral regions of scales in Nemipterus species suggests a functional role in reducing frictional forces through mechanical anchoring, a mechanism previously described in several other fish taxa12,69,70. In addition, the circuli of the rostral field bear a higher density of denticles compared to those of the lateral fields, which likely enhances the attachment of scales to the underlying skin9. Yang, et al.71 proposed that the circuli and other microroughness features on the surface of fish scales serve to concentrate tensile stresses within the depressions between circuli during scale bending. By restricting tensile stress to these regions, the overall stress experienced during bending is reduced, thereby influencing the magnitude of force required to achieve a given level of deformation62.The present study observed that the outer lateral circuli of N. japonicus, as well as the inner and outer lateral circuli of scales in specific locations of the other two Nemipterus species examined, lacked denticles. The absence of denticles in certain regions of the scales suggest that anchoring is not crucial in these regions, particularly in newly formed circuli3. These findings are consistent with the observations of Mekkawy, et al.12, Mekkawy, et al.3 and Mahmoud, et al.72.In the studied nemipterid species, the focus exhibits an oval or circular shape, encircled by elongated ridges that gradually diminish toward the caudal field, and is typically positioned in the posterior region of the scale. This observation aligns with the findings of Devi, et al.7 for N. japonicus. Although the position of the scale focus can vary among fish species18, it is generally considered stable once formed and does not shift throughout the lifespan of the individual73. Teimori, et al.74 further indicated that the shape and size of the focus are influenced by a combination of genetic factors, environmental conditions, and developmental processes. Variation in focus morphology has been widely reported among different fish taxa, including triangular, rectangular, circular, and oval forms1,3,56,75.The caudal region of Nemipterus scales lacks conventional microstructural features such as circuli and grooves; instead, it is distinguished by ornamentations composed of granulations (tubercles) and ctenii. Similar structural patterns have been documented in previous studies3,7,15,45,51. Additionally, the caudal field of the scales in the studied nemipterid species contained multiple pores, which likely function as minute canaliculi facilitating mucus secretion. This mucus layer may play an essential role in wound healing and in reducing hydrodynamic drag at the fish water interface, thereby enhancing locomotory efficiency76. Also, Teimori, et al.77 suggested that modification in the ornamentation of the posterior region may have hydrodynamic significance, and such features are subject to modification throughout ontogenetic development.Previous studies have demonstrated the taxonomic significance of lateral line scales in fish classification51,78,79,80. Khalil, et al.81 reported substantial genus-level variability in the anterior–posterior placement of the lateral line tube in Brycinus nurse and Alestes baremose. In contrast, the Nemipterus species examined in the present study showed no distinct species-specific pattern in lateral line scale morphology.In this investigation, the lateral line canal ran parallel to the anteroposterior axis of the scale, with posterior and anterior openings positioned medially within the focus region and separated from their respective scale margins. These observations correspond with the findings of Voronina and Hughes78, who examined more than 1,000 teleost species, including N. zysron, and classified such scales as Unmodified Tubular-Scalar I. In this type, the canal tube forms as an elevated structure on a fully developed elasmoid scale, with the tube and scale following independent developmental pathways82,83. The position of canal openings is considered an important diagnostic feature for differentiating taxa6,11,79.In general, Fish scale morphology has been shown to be most effective for distinguishing species within the same family that exhibit differing body morphologies, moderately effective for differentiation within fish families, and least effective for separating distantly related taxa17. Variations in scale shape and structure are influenced by multiple environmental factors, including temperature, water currents, population density, food availability, and other ecological conditions59,84. In addition to environmental influences, genetic factors have also been highlighted as critical determinants of the taxonomic utility of scale characteristics85.ConclusionsThis study demonstrates that scale morphology, together with traditional morphometric and geometric shape analyses, provides powerful tools for interspecific discrimination among Nemipterus species. The observed differences in scale characteristics, including surface ornamentation and structural features, underscore the value of scale-based assessments for species differentiation and stock identification. Collectively, these findings highlight the promising utility of scale morphology as a cost-effective approach for refining taxonomic classification and supporting conservation efforts. Nevertheless, classification accuracy may be enhanced by incorporating additional scale measurements and traits, such as scale thickness, flexibility, circuli counts, and lateral line canal patterns. Further comparative studies are required to validate the potential use of scale morphology and architectural design in species identification and phylogenetic applications within Nemipterus and related taxa.

    Data availability

    The data supporting this article are available from the corresponding author upon reasonable request.
    AbbreviationsBDFS:
    Beneath the anterior portion of the dorsal fin scales
    POS:
    Post-operculum scales
    BLLS:
    Beneath the lateral line, between the pectoral and pelvic fins scales
    CPS:
    Caudal peduncle directly above the lateral line scales
    ALLS:
    Anterior lateral line scales
    MLLS:
    Middle lateral line scales
    PLLS:
    Posterior lateral line scales
    R1:
    Primary radii
    R2:
    Secondary radii
    L:
    Scale length
    L1:
    Rostral field length
    L2:
    Caudal field length
    W:
    Scale width
    ReferencesAl Jufaili, S. M., Echreshavi, S. & Esmaeili, H. R. Scales surface topography: Comparative ultrastructural and decorative characteristics of a modern elasmoid fish scales in a cyprinid fish, Garra Shamal (Teleostei: Cyprinidae) using digital optical light and scanning electron microscope imaging. Microsc. Res. Tech. 86, 97–114. https://doi.org/10.1002/jemt.24263 (2023).Article 
    PubMed 

    Google Scholar 
    Viertler, A., Salzburger, W. & Ronco, F. Comparative scale morphology in the adaptive radiation of cichlid fishes (Perciformes: Cichlidae) from lake Tanganyika. Biol. J. Linn. Soc. 134, 541–556. https://doi.org/10.1093/biolinnean/blab099 (2021).Article 

    Google Scholar 
    Mekkawy, I. A. A., Mahmoud, U. M., El-Mahdy, S. M. & Essa, F. Morphometrics, geometrics and microstructures of scales of three fish species of genus Gerres from the Red Sea, Egypt. Fish. Sci. 91, 217–235. https://doi.org/10.1007/s12562-024-01838-2 (2025).Article 

    Google Scholar 
    Pacheco-Almanzar, E., Loza-Estrada, N. & Ibáñez, A. L. Do the fish scales shape of Mugil Curema reflect the genetic structure using microsatellites markers and the Mexican marine ecoregions classification? Front. Mar. Sci. 7 https://doi.org/10.3389/fmars.2020.00166 (2020).Purrafee Dizaj, L., Esmaeili, H. R., Teimori, A. & Abbasi, K. Comparative microscopic examination of scales in 21 Clupeid species from the Caspian sea and the Indo-Pacific regions. Micron 137, 102911. https://doi.org/10.1016/j.micron.2020.102911 (2020).Article 
    PubMed 

    Google Scholar 
    Matondo, D. A. P., Torres, M. A. J., Tabugo, S. R. M. & Demayo, C. G. Describing variations in scales between sexes of the yellowstriped goatfish, Upeneus vittatus (Forskål, 1775) (Perciformes: Mullidae). Egypt. Acad. J. Biol. Sci. (B- Zool.). 2, 37–50. https://doi.org/10.21608/eajbsz.2010.15911 (2010).Article 

    Google Scholar 
    Devi, P. S. S., Leemon, N., Suresh, S., Kumar, S. S. & Panicker, S. P. Characterization of the scale morphology and salient features of the selected marine fishes. Uttar Pradesh J. Zool. 42, 203–212 (2021).
    Google Scholar 
    Al Jufaili, S. M., Echreshavi, S., Esmaeili, H. R. & Al Alawi, M. K. Scales and otoliths as identity cards of the Indian oil sardine sardinella longiceps (Teleostei: Clupeiformes) populations: Ultrastructure and ornamentation characteristics using light and scanning electron microscopy. Acta Zool. 104, 380–397. https://doi.org/10.1111/azo.12418 (2022).Article 

    Google Scholar 
    Al Jufaili, S. M., Masoumi, A. H., Esmaeili, H. R., Jawad, L. & Teimori, A. Morphological and microstructural characteristics of scales in longnose goby Awaous Jayakari (Teleostei: Gobiidae): light and scanning electron microscopy approaches. Microsc. Res. Tech. 84, 3128–3149. https://doi.org/10.1002/jemt.23871 (2021).Article 
    PubMed 

    Google Scholar 
    Esmaeili, H. R., Zarei, F., Sanjarani Vahed, N. & Masoudi, M. Scale morphology and phylogenetic character mapping of scale-surface microstructures in sixteen Aphanius species (Teleostei: Aphaniidae). Micron 119, 39–53. https://doi.org/10.1016/j.micron.2019.01.002 (2019).Article 
    PubMed 

    Google Scholar 
    Harabawy, A. S. A., Mekkawy, I. A. A. & Alkaladi, A. Identification of three fish species of genus Plectorhynchus from the Red Sea by their scale characteristics. Life Sci. J. 9, 4472–4485 (2012).
    Google Scholar 
    Mekkawy, I. A. A., Wassif, E. T. & Basmidi, A. A. M. Scale characteristics of three Lutjanus species (Family: Lutjanidae) from the Red Sea, Egypt. J. Fish. Aquat. Sci. 6, 506–522. https://doi.org/10.3923/jfas.2011.506.522 (2011).Article 

    Google Scholar 
    Sadeghi, R., Esmaeili, H. R., Teimori, A., Ebrahimi, M. & Gholamhosseini, A. Comparative ultrastructure and ornamentation characteristics of scales in gobiid species (Teleostei: Gobiidae) using the scanning electron microscope. Microsc. Res. Tech. 84, 1243–1256. https://doi.org/10.1002/jemt.23683 (2021).Article 
    PubMed 

    Google Scholar 
    Mekkawy, I. A. A., Harabawy, A. S. A., Adam, E. A. & Mohamed, A. I. Biology and population dynamics of Lates niloticus (Linnaeus, 1758) (Family Centropomidae) from lake Nasser, Egypt. J. Egypt. Soc. Biotechnol. Environ. Sci. 10, 191–234 (2007).
    Google Scholar 
    Echreshavi, S., Esmaeili, H. R., Teimori, A., Safaie, M. & Owfi, F. Hidden morphological and structural characteristics in scales of mullid species (Teleostei: Mullidae) using light and scanning electron digital imaging. Microsc. Res. Tech. 84, 2749–2773. https://doi.org/10.1002/jemt.23837 (2021).Article 
    PubMed 

    Google Scholar 
    Ibañez, A. L., Cowx, I. G. & O’Higgins, P. Geometric morphometric analysis of fish scales for identifying genera, species, and local populations within the Mugilidae. Can. J. Fish. Aquat. Sci. 64, 1091–1100. https://doi.org/10.1139/f07-075 (2007).Article 

    Google Scholar 
    Ibáñez, A. L., Jawad, L. A., David, B., Rowe, D. & Ünlü, E. The morphometry of fish scales collected from new Zealand and Turkey. N. Z. J. Zool. 50, 318–328. https://doi.org/10.1080/03014223.2022.2035413 (2022).Article 

    Google Scholar 
    Sabbah, N., Teimori, A. & Hesni, M. A. Digital light microscopy to characterize the scales of two goatfishes (Perciformes; Mullidae). Microsc. Res. Tech. 84, 180–191. https://doi.org/10.1002/jemt.23576 (2021).Article 
    PubMed 

    Google Scholar 
    Șerban, C., & Grigoraş, G. Structural and morphometric study of scales in Petzea Rudd (Scardinius Racovitzai Müller 1958). Appl. Ecol. Environ. Res. 16, 6063–6076. https://doi.org/10.15666/aeer/1605_60636076 (2018).Article 

    Google Scholar 
    Traverso, F. et al. New insights into geometric morphometry applied to fish scales for species identification. Animals 14, 1090 (2024).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fricke, R., Eschmeyer, W. N. & Van der Laan, R. N. (eds) (ed California Academy of Sciences) (San Francisco, 2025).Russell, B. C. A review of the threadfin Breams of genus Nemipterus (Nemipteridae) from Japan and Taiwan, with description of new species. Jpn. J. Ichthyol. 39, 295–310. https://doi.org/10.11369/jji1950.39.295 (1993).Article 

    Google Scholar 
    Russell, B. C. FAO Species Catalogue. Vol.12. Family Nemipteridae. An Annotated and Illustrated Catalogue of Nemipterid Species Known to Date, Vol. 12 (FAO, 1990).Karuppasamy, K., Kingston, S. D., Jawahar, P. & Vidhya, V. Spatio-temporal variation in the diversity of threadfin Breams (Family: Nemipteridae) from Wadge Bank, South India. J. Entomol. Zool. Stud. 6, 450–454 (2018).
    Google Scholar 
    Akgun, Y. & Akoglu, E. Randall’s threadfin Bream (Nemipterus randalli, Russell 1986) poses a potential threat to the Northeastern mediterranean sea food web. Fishes 8, 402. https://doi.org/10.3390/fishes8080402 (2023).Article 

    Google Scholar 
    Demirci, S., Demirci, A. & Şimşek, E. The validation of different growth models of randall’s threadfin Bream, Nemipterus randalli (Russell, 1986), in Northeastern mediterranean sea. Pak J. Zool. 52, 1113–1119. https://doi.org/10.17582/journal.pjz/20180327130349 (2020).Article 

    Google Scholar 
    El-Haweet, A. E. A. Biological studies of the invasive species Nemipterus japonicus (Bloch, 1791) as a red sea immigrant into the mediterranean. Egypt. J. Aquat. Res. 39, 267–274 (2013).Article 

    Google Scholar 
    Lisamy, S. E. A. et al. Reproductive aspects of the Japanese threadfin bream, nemipterus japonicus (Bloch, 1791) in the Southern Java waters (FMA-RI 573). E3S Web Conf. 442, 1–13. https://doi.org/10.1051/e3sconf/202344201025 (2023).Article 

    Google Scholar 
    Mehanna, S. F. & Farouk, A. E. Length-Weight relationship of 60 fish species from the Eastern mediterranean Sea, Egypt (GFCM-GSA 26). Front. mar. sci. 8, 1–7. https://doi.org/10.3389/fmars.2021.625422 (2021).Article 

    Google Scholar 
    El-Ganainy, A. A., Khalil, M. T., El-Bokhty, E. E. E., Saber, M. A., & Abd El-Rahman, F. A. A. Assessment of three nemipterid stocks based on trawl surveys in the Gulf of Suez, red sea. Egypt. J. Aquat. Res. 44, 45–49. https://doi.org/10.1016/j.ejar.2018.02.005 (2018).Article 

    Google Scholar 
    Joshi, K. K. Population dynamics of Nemipterus japonicus (Blotch) trawling grounds off Cochin. Indian J. Fish. 57, 7–12 (2010).
    Google Scholar 
    Özen, M. R. & Çetİnkaya, O. Population composition, growth and fisheries of Nemipterus randalli Russell, 1986 in Antalya Gulf, mediterranean Sea, Turkey. Act. Aqua Tr. 16, 330–337. https://doi.org/10.22392/actaquatr.681309 (2020).Article 

    Google Scholar 
    Tonie, N., Idris, M. H., Al-Asif, A., Hussin, W. M. R. W. & Kamal, A. M. Population characteristics of the Japanese threadfin Bream Nemipterus japonicus (Bloch, 1791) (Actinopterygii: Nemipteridae) at Bintulu coast, Sarawak, South China sea. Acta Zool. Bulg. 75, 273–283. https://doi.org/10.6084/m9.figshare.23614788.v1 (2023).Article 

    Google Scholar 
    Naqash, S. Y. & Nazeer, R. A. Antioxidant activity of hydrolysates and peptide fractions of Nemipterus japonicus and Exocoetus volitans muscle. J. Aquat. Food Prod. Technol. 19, 180–192. https://doi.org/10.1080/10498850.2010.506256 (2010).Article 

    Google Scholar 
    Ay, İ., Çiftçi, N. & Ayas, D. Morphometric comparison of different populations of Nemipterus randalli Russell 1986 distributed in the mediterranean Coasts of Turkey. Adv. Underw. Sci. J. 2, 1–8 (2021).
    Google Scholar 
    Ning, P., Sha, Z., Hebert, P. D. N. & Russell, B. The taxonomic status of Japanese threadfin Bream Nemipterus japonicus (Bloch, 1791) (Perciformes: Nemipteridae) with a redescription of this species from the South China sea based on morphology and DNA barcodes. J. Ocean. Univ. Chin. Ocean. Coast Sea Res. 14, 178–184. https://doi.org/10.1007/s11802-015-2609-x (2015).Article 

    Google Scholar 
    Ogwang, J., Bariche, M. & Bos, A. R. Genetic diversity and phylogenetic relationships of threadfin Breams (Nemipterus spp.) from the Red Sea and Eastern mediterranean sea. Genome 64, 207–216. https://doi.org/10.1139/gen-2019-0163/M (2021).Article 
    PubMed 

    Google Scholar 
    Öztürk, B. Status of alien species in the black and mediterranean seas, in Studies and Reviews. General Fisheries Commission for the Mediterranean, Vol. 87 (FAO, (2010).Aydın, İ. & Akyol, O. Occurrence of nemipterus randalli Russell, 1986 (Nemipteridae) off Izmir Bay, Turkey. J. Appl. Ichthyol. 33, 533–534. https://doi.org/10.1111/jai.13331 (2017).Article 

    Google Scholar 
    Fisher, W. & Whitehead, P. J. P. FAO Species Identification Sheets for Fisheries Purposes. Eastern Indian Ocean (fishing area 57) and Western Central Pacific (fishing area 71), Vol. 3 (FAO, 1974).Lelli, S., Colloca, F., Carpentieri, P. & Russell, B. C. The threadfin Bream Nemipterus randalli (Perciformes: Nemipteridae) in the Eastern mediterranean sea. J. Fish. Biol. 73, 740–745. https://doi.org/10.1111/j.1095-8649.2008.01962.x (2008).Article 

    Google Scholar 
    Russ, J. C. Computer-Assisted Microscopy, 1 edn (Springer, 1990).Saber, M. A. Size selection by diamond and square mesh codends for the most commercial demersal fishes in the Gulf of Suez Ph.D thesis, Ain Shams (2017).tpsUtil File Utility Program v. Version 1.58 (Department of Ecology and Evolution (State University of New York, 2010).Hassanien, E. M. Biological Studies, Stock Assessment and Fisheries Management of Some Species of Family Mullidae from the Mediterranean Sea at Alexandria Ph.D. thesis, Assiut University (2017).IBM SPSS Statistics for Windows. Version 22.0 (Armonk (IBM Corp, 2019).Hammer, Ø., Harper, D. A. & Ryan, P. D. PAST: Paleontological statistics software package for education and data analysis. Palaeontol. Electron. 4, 9 (2001).
    Google Scholar 
    Klingenberg, C. P. MorphoJ: An integrated software package for geometric morphometrics. Mol. Ecol. Resour. 11, 353–357. https://doi.org/10.1111/j.1755-0998.2010.02924.x (2011). https://doi.org/https://Article 
    PubMed 

    Google Scholar 
    Rohlf, F. J. & Slice, D. Extensions of the procrustes method for the optimal superimposition of landmarks. Syst. Biol. 39, 40–59. https://doi.org/10.2307/2992207 (1990).Article 

    Google Scholar 
    Cavalcanti, M., Monteiro, L. & Lopes, P. Landmark-based morphometric analysis in selected species of serranid fishes (Perciformes: Teleostei). Zool. Stud. 38, 287–224 (1999).
    Google Scholar 
    Renjith, R. K., Jaiswar, A. K., Chakraborty, S. K., Jahageerdar, S. & Sreekanth, G. B. Application of scale shape variation in fish systematics-an illustration using six species of the family Nemipteridae (Teleostei: Perciformes). Indian J. Fish. 61, 88–92 (2014).
    Google Scholar 
    Teimori, A. The scale characteristics of two Aphanius species from Southern Iran (Teleostei: Aphaniidae). Zool. Middle East. 64, 219–227. https://doi.org/10.1080/09397140.2018.1475120 (2018).Article 

    Google Scholar 
    Ujjania, N. & Jaiswar, A. Scale morphology an additional tool for taxonomy and fish identification with reference to Nemipteridae fishes (N. japonicus, N. bipunctatus and N. randalli). Flora Fauna. 30, 343–351. https://doi.org/10.33451/florafauna.v30i2pp343-351 (2024).Article 

    Google Scholar 
    Spinner, M. et al. Mechanical behavior of ctenoid scales: Joint-like structures control the deformability of the scales in the flatfish Solea Solea (Pleuronectiformes). Acta Biomater. 92, 305–314. https://doi.org/10.1016/j.actbio.2019.05.011 (2019).Article 
    PubMed 

    Google Scholar 
    Ibañez, A. L., Cowx, I. G. & O’higgins, P. Variation in elasmoid fish scale patterns is informative with regard to taxon and swimming mode. Zool. J. Linn. Soc. 155, 834–844. https://doi.org/10.1111/j.1096-3642.2008.00465.x (2009).Article 

    Google Scholar 
    Mohammad, A. S. Population Dynamics and Stock Assessment of Some Species of Genus Cephalopholis and Genus Variola from the Red Sea, Egypt M.Sc. thesis, Assiut University (2007).Jawad, L. A. & Al-Jufaili, S. M. Scale morphology of greater Lizardfish Saurida tumbil (Bloch, 1795) (Pisces: Synodontidae). J. Fish. Biol. 70, 1185–1212. https://doi.org/10.1111/j.1095-8649.2007.01385.x (2007).Article 

    Google Scholar 
    El-Mahdy, S. M. Biological Studies, Population Dynamics and Stock Assessment of Acanthopagrus Bifasciatus (Forsskal, 1775) from the Red Sea, Egypt Ph.D. thesis, Assiut University, (2018).Raffealla, N. & Nath, B. R. Comparative study of fish scale using scanning electron microscopy in two cyprinid fishes (Neolissochilus Hexagonolepis and Neolissochilus hexastichus) found in Meghalaya, North-East India. Int. J. Life Sci. 8, 77–82 (2020).
    Google Scholar 
    Alkaladi, A., Harabawy, A. S. & Mekkawy, I. A. Scale characteristics of two fish species, Acanthopagrus bifasciatus (Forsskål, 1775) and Rhabdosargus Sarba (Forsskål, 1775) from the red sea at Jeddah, Saudi Arabia. Pak. J. Biol. Sci. 16, 362–371. https://doi.org/10.3923/pjbs.2013.362.371 (2013).Article 
    PubMed 

    Google Scholar 
    Esmaeili, H. R. & Gholami, Z. Scanning electron microscopy of the scale morphology in cyprinid fish, Rutilus frisii Kutum Kamenskii, 1901 (Actinopterygii: cyprinidae). Iran. J. Fish. Sci. 10, 155–166 (2011).
    Google Scholar 
    Wainwright, D. K. & Lauder, G. V. Three-dimensional analysis of scale morphology in Bluegill sunfish, lepomis macrochirus. Zoology 119, 182–195. https://doi.org/10.1016/j.zool.2016.02.006 (2016). https://doi.org/https://doi.Article 
    PubMed 

    Google Scholar 
    Farah-Ayuni, F., Muse, A. O., Samat, A. & Shukor, M. N. Comparative scale morphologies in common freshwater fishes of Peninsular Malaysia – A case study. AIP Conf. Proc. 1784 https://doi.org/10.1063/1.4966850 (2016).Farinordin, F. A., Nilam, W. S. W., Husin, S. M., Samat, A. & Nor, S. M. Scale morphologies of freshwater fishes at Tembat forest Reserve, Terengganu, Malaysia. Sains Malays. 46, 1429–1439. https://doi.org/10.17576/jsm-2017-4609-11 (2017). https://doi.org/http:Article 

    Google Scholar 
    Mekkawy, I. A. A. & Abdel-Rahman, G. H. Comparative study of scales of five species of parrotfishes (Family Scaridae) from the red Sea, Hurghada, Egypt with emphasis on their functional morphology. Egypt. J. Zool. 44, 545–580 (2005).
    Google Scholar 
    Lippitsch, E. Scale morphology and squamation patterns in cichlids (Teleostei, Perciformes): A comparative study. J. Fish. Biol. 37, 265–291. https://doi.org/10.1111/j.1095-8649.1990.tb05858.x (1990). Article 

    Google Scholar 
    Gholami, Z., Teimori, A., Esmaeili, H. R., Schulz-Mirbach, T. & Reichenbacher, B. Scale surface microstructure and scale size in the tooth-carp genus Aphanius (Teleostei, Cyprinodontidae) from endorheic basins in Southwest Iran. Zootaxa 3619, 467–490. https://doi.org/10.11646/zootaxa.3619.4.5 (2013).Article 
    PubMed 

    Google Scholar 
    Ferrito, V., Pappalardo, A. M., Fruciano, C. & Tigano, C. Morphology of scale lepidonts in the genus Aphanius (Teleostei, Cyprinodontidae) using SEM. Ital. J. Zool. 76, 173–178. https://doi.org/10.1080/11250000802555684 (2009).Article 

    Google Scholar 
    Esmaeili, H. R. et al. Scale surface microstructure and scale size in three Mugilid fishes (Teleostei, Mugilidae) of Iran from three different habitats. IUFS J. Biol. 73, 31–42. https://doi.org/10.18478/IUFSJB.62393 (2014).Article 

    Google Scholar 
    Zahid, H. et al. Scale surface structure of Mugil cephalus (Teleostei; Mugilidae) using scanning electron microscopy (SEM). Biol. Forum Int. J. 7, 1–4 (2015).
    Google Scholar 
    Yang, W., Chen, I. H., McKittrick, J. & Meyers, M. A. Flexible dermal armor in nature. JOM 64, 475–485. https://doi.org/10.1007/s11837-012-0301-9 (2012).Article 

    Google Scholar 
    Mahmoud, U. M., El-Gammal, F. I., Mehanna, S. F. & El-Mahdy, S. M. Scale characteristics of Acanthopagrus bifasciatus (Forsskål, 1775) from the Southern red Sea, Egypt. Int. J. Fish. Aquat. 5, 417–422 (2017).
    Google Scholar 
    Ganzon, M. A. M., Torres, M. A., Gorospe, J. J. & Demayo, C. G. in 2nd International Conference on Environment and BioScience, IPCBEE 44.Teimori, A., Iranmanesh, N., Hesni, M. A. & Motamedi, M. Microanalysis of scale morphology in killifish, Aphaniops hormuzensis inhabiting ecologically diverse environments (Cyprinodontiformes; Aphaniidae). Micron 140, 102949. https://doi.org/10.1016/j.micron.2020.102949 (2021a).Article 
    PubMed 

    Google Scholar 
    Helfman, G. S., Collette, B. B., Facey, D. E. & Bowen, B. W. The Diversity of Fishes: Biology, Evolution, and Ecology, Second edn (Blackwell, 2009).Mahmoud, U. M., Mekkawy, I. A. A. & Harabawy, A. S. A. Scale characteristics of seven species of genus Lethrinus (Family Lethrinidae) from the Red Sea, Egypt. Egypt. J. Zool. 44, 545–580 (2005).
    Google Scholar 
    Teimori, A., Motamedi, M., Amiri, V. & Hesni, M. A. Ultramicroscopy of structures involved in the posterior region of scales in two Flathead fishes (Teleostei: Perciformes). Int. J. Aquat. Biol. 9, 15–22 (2021b).
    Google Scholar 
    Voronina, E. P. & Hughes, D. R. Lateral line scale types and review of their taxonomic distribution. Acta Zool. 99, 65–86. https://doi.org/10.1111/azo.12193 (2018).Article 

    Google Scholar 
    Yu, H., Wang, Z., Qin, H. & Chen, Y. An automatic detection and counting method for fish lateral line scales of underwater fish based on improved YOLOv5. IEEE Access. 11, 143616–143627. https://doi.org/10.1109/access.2023.3343429 (2023).Article 

    Google Scholar 
    Webb, J. F. & Ramsay, J. B. New interpretation of the 3-D configuration of lateral line scales and the lateral line Canal contained within them. Copeia 105, 339–347 (2017).Article 

    Google Scholar 
    Khalil, A., Yoakim, E. G. & Mekkawy, I. A. A. Identification of two nile fish species of the genus Alestes by scale characteristics. Bull. Fac. Sci. Assiut Univ. E Zool. 11, 185–207. https://doi.org/10.21608/avmj.1983.191492 (1982).Article 

    Google Scholar 
    Voronina, E. P. & Hughes, D. R. Types and development pathways of lateral line scales in some teleost species. Acta Zool. 94, 154–166. https://doi.org/10.1111/j.1463-6395.2011.00534.x (2013). https://doi.org/.Article 

    Google Scholar 
    Wonsettler, A. L. & Webb, J. F. Morphology and development of the multiple lateral line canals on the trunk in two species of hexagrammos (Scorpaeniformes, Hexagrammidae). J. Morphol. 233, 195–214. https://doi.org/10.1002/(SICI)1097-4687(199709)233:3%3C195::AID-JMOR1%3E3.0.CO;2-3 (1997). https://doi.org/https://doi.org/3.0.CO;2-3″ data-track-item_id=”10.1002/(SICI)1097-4687(199709)233:3<195::AID-JMOR1>3.0.CO;2-3″ data-track-value=”article reference” data-track-action=”article reference” href=”https://doi.org/10.1002%2F%28SICI%291097-4687%28199709%29233%3A3%3C195%3A%3AAID-JMOR1%3E3.0.CO%3B2-3″ aria-label=”Article reference 83″ data-doi=”10.1002/(SICI)1097-4687(199709)233:3<195::AID-JMOR1>3.0.CO;2-3″>Article 
    PubMed 

    Google Scholar 
    Ibáñez, A. L., Pacheco-Almanzar, E. & Cowx, I. G. Does compensatory growth modify fish scale shape? Environ. Biol. Fishes. 94, 477–482. https://doi.org/10.1007/s10641-011-9962-4 (2012).Article 

    Google Scholar 
    Maheboob, S. I. Study of morphology of scale in three teleost species from Purna river basin in part of Parbhani Districts, Maharashtra India. Res. Rev. Int. J. Multidiscip. 6, 14–19. https://doi.org/10.31305/rrijm.2020.v06.i02.004 (2021).Article 

    Google Scholar 
    Download referencesAcknowledgementsThe authors thank the National Institute of Oceanography and Fisheries and the Electron Microscope Unit, Assiut University, Egypt for providing invaluable support and access to their facilities during the course of this study.FundingOpen access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB). The authors declare that no funds, grants or other support were received during the preparation of this manuscript.Author informationAuthors and AffiliationsZoology Department, Faculty of Science, Assiut University, Assiut, 71516, EgyptImam A. A. Mekkawy, Usama M. Mahmoud & Ola I. MuhammadFisheries Division, National Institute of Oceanography and Fisheries, NIOF, Hurghada, Red Sea, EgyptSamia M. El-MahdyAuthorsImam A. A. MekkawyView author publicationsSearch author on:PubMed Google ScholarUsama M. MahmoudView author publicationsSearch author on:PubMed Google ScholarSamia M. El-MahdyView author publicationsSearch author on:PubMed Google ScholarOla I. MuhammadView author publicationsSearch author on:PubMed Google ScholarContributionsI.A.A.M., U.M.M., S.M.E. and O.I.M. designed the study. S.M.E. and O.I.M. collected data and drafted the initial manuscript. I.A.A.M. and O.I.M. analysed data and contributed to the final version of the manuscript. I.A.A.M., U.M.M. and S.M.E. supervised the study. All authors provide consent for publication.Corresponding authorCorrespondence to
    Ola I. Muhammad.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Ethics approval and consent to participate
    All specimens in this study were collected from the fish market at Hurghada, Red Sea, Egypt, no fish were killed during this study. No ethical approval was required for this study. All the authors have provided their consent to participate in this study.

    Additional informationPublisher’s noteSpringer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Rights 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 articleMekkawy, I.A.A., Mahmoud, U.M., El-Mahdy, S.M. et al. Scale morphometry, geometry and ultrastructure of three Nemipterus species from the Egyptian part of the Red Sea.
    Sci Rep 15, 43871 (2025). https://doi.org/10.1038/s41598-025-30040-2Download citationReceived: 08 July 2025Accepted: 20 November 2025Published: 15 December 2025Version of record: 16 December 2025DOI: https://doi.org/10.1038/s41598-025-30040-2Share 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
    KeywordsNemipterusScale characteristicsScanning electron microscopeMorphometricsGeometricsFisheriesRed sea More

  • in

    Plant community data along elevational gradients in China’s 17 mountains

    AbstractMountains in China are crucial for biodiversity conservation due to unique topography and climate, providing essential habitats and refugia for many plant species. Standardized open datasets along elevational gradients across multiple mountains remain limited. Here we used standardized field protocols to collect plant diversity data of 370 permanent sampling plots along elevations in 17 mountains. Species identity and abundance of all woody plants with ≥ 1 cm diameter at breast height were recorded. We calculated species-level basal area, abundance, and relative importance value for all plants, and separately for two vegetation layers. The dataset spans 46° longitude, 24° latitude, ranges from 166 to 3,835 m a.s.l., and includes 1,493 species from 121 families and 450 genera. It covers nearly all major ecosystems from tropical rainforests to tundra, providing baseline data for studying plant diversity changes along elevations and latitudes. This dataset enables direct comparisons across mountains, helping evaluate impacts of climate and land-use changes on species range shifts and ecosystem transitions, and inform conservation strategies for mountain ecosystems.

    Similar content being viewed by others

    The China plant trait database version 2

    Article
    Open access
    15 December 2022

    Growth characteristics of Cunninghamia lanceolata in China

    Article
    Open access
    28 October 2022

    The mid-domain effect of mountainous plants is determined by community life form and family flora on the Loess Plateau of China

    Article
    Open access
    26 May 2021

    Data availability

    The dataset supporting this Data Descriptor has been deposited in Figshare and is publicly available at https://doi.org/10.6084/m9.figshare.30877097.
    Code availability

    No custom code was used to generate or process the data described in this study.
    ReferencesRahbek, C. et al. Humboldt’s enigma: what causes global patterns of mountain biodiversity? Science 365, 1108–1113 (2019).
    Google Scholar 
    Perrigo, A., Hoorn, C. & Antonelli, A. Why mountains matter for biodiversity. Journal of Biogeography 47, 315–325 (2020).
    Google Scholar 
    Hoorn, C., Mosbrugger, V., Mulch, A. & Antonelli, A. Biodiversity from mountain building. Nature Geoscience 6, 154–154 (2013).
    Google Scholar 
    Körner, C. & Spehn, E. M. Mountain biodiversity: a Global Assessment. (Parthenon, Boca Raton, 2002).
    Google Scholar 
    Sundqvist, M. K., Sanders, N. J. & Wardle, D. A. Community and ecosystem responses to elevational gradients: processes, mechanisms, and insights for global change. Annual Review of Ecology, Evolution, and Systematics 44, 261–280 (2013).
    Google Scholar 
    Muellner-Riehl, A. N. Mountains as evolutionary arenas: patterns, emerging approaches, paradigm shifts, and their implications for plant phylogeographic research in the Tibeto-Himalayan region. Frontiers in Plant Science 10 (2019).Peters, M. K. et al. Predictors of elevational biodiversity gradients change from single taxa to the multi-taxa community level. Nature Communications 7, 13736 (2016).
    Google Scholar 
    La Sorte, F. A. & Jetz, W. Projected range contractions of montane biodiversity under global warming. Proceedings of the Royal Society B: Biological Sciences 277, 3401–3410 (2010).
    Google Scholar 
    He, N. et al. Predicting ecosystem productivity based on plant community traits. Trends in Plant Science 28, 43–53 (2023).
    Google Scholar 
    Åkesson, A. et al. The importance of species interactions in eco-evolutionary community dynamics under climate change. Nature Communications 12, 4759 (2021).
    Google Scholar 
    Vellend, M. The Theory of Ecological Communities (Princeton University Press, 2016).Ehrlén, J. & Morris, W. F. Predicting changes in the distribution and abundance of species under environmental change. Ecology Letters 18, 303–314 (2015).
    Google Scholar 
    Ohlmann, M. et al. Quantifying the overall effect of biotic interactions on species distributions along environmental gradients. Ecological Modelling 483, 110424 (2023).
    Google Scholar 
    Mi, X. et al. The global significance of biodiversity science in China: an overview. National Science Review 8, nwab032 (2021).
    Google Scholar 
    López-Pujol, J., Zhang, F.-M., Sun, H.-Q., Ying, T.-S. & Ge, S. Mountains of southern China as “Plant museums” and “Plant cradles”: evolutionary and conservation insights. Mountain Research and Development 31, 261–269 (2011).
    Google Scholar 
    Tang, Z., Wang, Z., Zheng, C. & Fang, J. Biodiversity in China’s mountains. Frontiers in Ecology and the Environment 4, 347–352 (2006).
    Google Scholar 
    Zhang, Z., He, J.-S., Li, J. & Tang, Z. Distribution and conservation of threatened plants in China. Biological Conservation 192, 454–460 (2015).
    Google Scholar 
    Condit, R. Tropical Forest Census Plots: Methods and Results from Barro Colorado Island, Panama and a Comparison with Other Plots (Springer, 1998).Karger, D. N. et al. CHELSA-W5E5: daily 1 km meteorological forcing data for climate impact studies. Earth System Science Data 15, 2445–2464 (2023).
    Google Scholar 
    WFO (World Flora Online). An online flora of all known plants. http://www.worldfloraonline.org (2025).Wu, Z., Raven, P. H. & Hong, D. Flora of China (Science Press, Beijing, 2005).Zhang, J. & Qian, H. U. Taxonstand: an R package for standardizing scientific names of plants and animals. Plant Diversity 45, 1–5 (2023).
    Google Scholar 
    Zhang, J., Qian, H. & Wang, X. An online version and some updates of R package U.Taxonstand for standardizing scientific names in plant and animal species. Plant Diversity 47, 166–168 (2025).
    Google Scholar 
    Jin, Y. & Qian, H. V. PhyloMaker2: an updated and enlarged R package that can generate very large phylogenies for vascular plants. Plant Diversity 44, 335–339 (2022).
    Google Scholar 
    Smith, S. A. & Brown, J. W. Constructing a broadly inclusive seed plant phylogeny. American Journal of Botany 105, 302–314 (2018).
    Google Scholar 
    Zanne, A. E. et al. Three keys to the radiation of angiosperms into freezing environments. Nature 506, 89–92 (2014).
    Google Scholar 
    Lembrechts, J. J. et al. SoilTemp: a global database of near-surface temperature. Global Change Biology 26, 6616–6629 (2020).
    Google Scholar 
    Wang, X. et al. Plant community data along elevational gradients in China’s 17 mountains, Figshare., https://doi.org/10.6084/m9.figshare.30877097 (2025).Ding, Y., Zang, R., Lu, X., Huang, J. & Xu, Y. The effect of environmental filtering on variation in functional diversity along a tropical elevational gradient. Journal of Vegetation Science 30, 973–983 (2019).
    Google Scholar 
    Huo, C., Zhang, Z., Hu, G. & Luo, Y. Altitude-related variation in carbon, nitrogen, and phosphorus contents and their stoichiometry of woody organs in the subtropical mountain forests, south China. Ecology and Evolution 15, e71451 (2025).
    Google Scholar 
    Wang, R. et al. Inconsistent elevational patterns of soil microbial biomass, diversity, and community structure on four elevational transects from subtropical forests. Applied Soil Ecology 201, 105462 (2024).
    Google Scholar 
    Shen, Y. et al. Mixed-species bird flocks re-assemble interspecific associations across an elevational gradient. Proceedings of the Royal Society B: Biological Sciences 289, 20221840 (2022).
    Google Scholar 
    Song, J. et al. Leaf and root traits show contrasting resource exploitation strategies, but converge along elevation in the Hengduan Mountain forests. Journal of Biogeography 52, e15157 (2025).
    Google Scholar 
    Chen, B. et al. What control home‐field advantage of foliar litter decomposition along an elevational gradient in subtropical forests? Plant and Soil, https://doi.org/10.1007/s11104-024-07165-w (2025).Luo, Y.-H. et al. Greater than the sum of the parts: how the species composition in different forest strata influence ecosystem function. Ecology Letters 22, 1449–1461 (2019).
    Google Scholar 
    Ma, L. et al. When microclimates meet soil microbes: temperature controls soil microbial diversity along an elevational gradient in subtropical forests. Soil Biology and Biochemistry 166, 108566 (2022).
    Google Scholar 
    Qian, S. et al. Conservation and development in conflict: regeneration of wild Davidia involucrata (Nyssaceae) communities weakened by bamboo management in south-central China. Oryx 52, 442–451 (2018).
    Google Scholar 
    Wang, J. et al. Spatial distribution pattern and correlation of dominant species in the arbor layer at a 25 hm2 forest plot in Lushan Mountain, China. Chinese Journal of Applied Ecology 34, 1491–1499 (2023).
    Google Scholar 
    Zhang, R. et al. A taxonomic and phylogenetic perspective on plant community assembly along an elevational gradient in subtropical forests. Journal of Plant Ecology 14, 702–716 (2021).
    Google Scholar 
    Wu, D. et al. Tree height and not climate influences intraspecific variations in wood parenchyma fractions of angiosperm species in a mountain forest of eastern China. American Journal of Botany 112, e70035 (2025).
    Google Scholar 
    Jia, S. et al. Neighbouring tree effects on leaf herbivory: insect specialisation matters more than host plant leaf traits. Journal of Ecology 112, 189–199 (2024).
    Google Scholar 
    Li, X. et al. Distinct strategies of soil bacterial generalists and specialists in temperate deciduous broad-leaved forests. Applied and Environmental Microbiology 91, e00992–25 (2025).
    Google Scholar 
    Wang, X. et al. Tree mycorrhizal associations strongly mediate soil microbial β-diversity along an elevational gradient in a warm-temperate forest. Applied Soil Ecology 205, 105776 (2025).
    Google Scholar 
    Fang, S. et al. Disturbance history, neighborhood crowding and soil conditions jointly shape tree growth in temperate forests. Oecologia 205, 295–306 (2024).
    Google Scholar 
    Tian, Z. et al. Wild apples are not that wild: conservation status and potential threats of Malus sieversii in the mountains of Central Asia biodiversity hotspot. Diversity 14, 489 (2022).
    Google Scholar 
    Download referencesAcknowledgementsWe are grateful to all the field crews in 17 mountains to collect the data, including Kankan Shang, Lin Chen, Qingni Song, Mingshui Zhao, Xin Wang, Yuzhuo Wang, Ran Zhang, Jiaxin Kong, Xianyu Yang, Oukai Zhang, Xuan Lv, Jiale Chen, Yaoshun Lu, Hongwei Zhang, Luwen Ma, Li Shu, Pengcheng Liu, Fang Wang, Xiaofan Shang, Jingchao Zhao, Junhong Chen, Mufan Sun, Min Guan, Pu Zheng, Yuetong Wang, Li Huang and Xijin Zhang. This work is part of the BEST (Biodiversity along Elevational Gradients: Shifts and Transitions; https://BEST-mountains.org) research network. This work was supported by the Fund of CAS Key Laboratory of Forest Ecology and Silviculture, Institute of Applied Ecology, Chinese Academy of Sciences (KLFES-2036, KLFES-2027), the National Natural Science Foundation of China (31500355, 31670630, 32071652, 32101280, 32230067, 32271616, 32301401, 32401334, 32471623, 32471852, and 41671047), the Innovation Program of Shanghai Municipal Education Commission (2023ZKZD36), the Jiangxi Natural Science Foundation (20242BAB25345, to Zhaochen Zhang), the Special Funding for Guangxi Bagui Young Top Scholar (to Zhonghua Zhang), the Excellent Young Scientist Program of Liaoning province (2024JH3/10200024), and the Doctoral Start-up Foundation of Liaoning Province (2024010292-JH3/101).Author informationAuthors and AffiliationsZhejiang Tiantong Forest Ecosystem National Observation and Research Station, School of Ecological and Environmental Sciences, East China Normal University, Shanghai, ChinaXiaoran Wang, Kun Song & Dingliang XingCollege of Life Sciences, Henan Agricultural University, Zhengzhou, ChinaYun Chen & Zhiliang YuanKey Laboratory of the Ministry of Education for Coastal and Wetland Ecosystems, College of the Environment and Ecology, Xiamen University, Xiamen, ChinaYuxin ChenKey Laboratory of Forest Ecology and Environment of the National Forestry and Grassland Administration, Ecology and Nature Conservation Institute, Chinese Academy of Forestry, Beijing, ChinaYi DingCAS Key Laboratory of Forest Ecology and Silviculture, Institute of Applied Ecology, Chinese Academy of Sciences, Shenyang, ChinaShuai Fang & Fei LinSchool of Ecology and Environment, Northwestern Polytechnical University, Xi’an, ChinaZhanqing Hao, Shihong Jia & Zuoqiang YuanKey Laboratory of Fujian Universities for Ecology and Resource Statistics, College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, ChinaZhongsheng HeKey Laboratory of Environment Change and Resources Use in Beibu Gulf, Ministry of Education, Nanning Normal University, Nanning, ChinaGang Hu & Zhonghua ZhangState Key Laboratory of Biocontrol, Innovation Center for Evolutionary Synthetic Biology, School of Life Sciences, Sun Yat-sen University, Guangzhou, ChinaBuhang Li & Jian ZhangCAS Key Laboratory for Plant Diversity and Biogeography of East Asia, Kunming Institute of Botany, Chinese Academy of Sciences, Kunming, ChinaJie Liu & Yahuang LuoGermplasm Bank of Wild Species in Southwest China, Kunming Institute of Botany, Chinese Academy of Sciences, Kunming, ChinaJie Liu & Yahuang LuoKey Laboratory of Agro-Forestry Environmental Processes and Ecological Regulation of Hainan Province, School of Ecology, Hainan University, Haikou, ChinaLan LiuLijiang Forest Biodiversity National Observation and Research Station, Kunming Institute of Botany, Chinese Academy of Sciences, Lijiang, ChinaYahuang LuoGuangxi Key Laboratory of Forest Ecology and Conservation, College of Forestry, Guangxi University, Nanning, ChinaYinghua LuoGuangdong Key Laboratory of Animal Conservation and Resource Utilization, Guangdong Public Laboratory of Wild Animal Conservation and Utilization, Institute of Zoology, Guangdong Academy of Sciences, Guangzhou, ChinaYong Shen & Qiang ZhangCollege of Forestry, Shanxi Agricultural University, Jinzhong, ChinaHoujuan Song & Xiuqing YangInstitute of Eco-Chongming, Shanghai, ChinaKun Song & Dingliang XingXinjiang Key Laboratory of Special Species Conservation and Regulatory Biology, College of Life Science, Xinjiang Normal University, Urumqi, ChinaZhongping TianJiangxi Provincial Key Laboratory for Bamboo Germplasm Resources and Utilization, Forestry College, Jiangxi Agricultural University, Nanchang, ChinaQingpei YangKey Laboratory of the Three Gorges Reservoir Region’s Eco-Environment, Ministry of Education, Chongqing University, Chongqing, ChinaYongchuan YangLushan Botanical Garden, Jiangxi Province and Chinese Academy of Sciences, Jiujiang, ChinaZhaochen ZhangAuthorsXiaoran WangView author publicationsSearch author on:PubMed Google ScholarYun ChenView author publicationsSearch author on:PubMed Google ScholarYuxin ChenView author publicationsSearch author on:PubMed Google ScholarYi DingView author publicationsSearch author on:PubMed Google ScholarShuai FangView author publicationsSearch author on:PubMed Google ScholarZhanqing HaoView author publicationsSearch author on:PubMed Google ScholarZhongsheng HeView author publicationsSearch author on:PubMed Google ScholarGang HuView author publicationsSearch author on:PubMed Google ScholarShihong JiaView author publicationsSearch author on:PubMed Google ScholarBuhang LiView author publicationsSearch author on:PubMed Google ScholarFei LinView author publicationsSearch author on:PubMed Google ScholarJie LiuView author publicationsSearch author on:PubMed Google ScholarLan LiuView author publicationsSearch author on:PubMed Google ScholarYahuang LuoView author publicationsSearch author on:PubMed Google ScholarYinghua LuoView author publicationsSearch author on:PubMed Google ScholarYong ShenView author publicationsSearch author on:PubMed Google ScholarHoujuan SongView author publicationsSearch author on:PubMed Google ScholarKun SongView author publicationsSearch author on:PubMed Google ScholarZhongping TianView author publicationsSearch author on:PubMed Google ScholarDingliang XingView author publicationsSearch author on:PubMed Google ScholarQingpei YangView author publicationsSearch author on:PubMed Google ScholarXiuqing YangView author publicationsSearch author on:PubMed Google ScholarYongchuan YangView author publicationsSearch author on:PubMed Google ScholarZhiliang YuanView author publicationsSearch author on:PubMed Google ScholarZuoqiang YuanView author publicationsSearch author on:PubMed Google ScholarQiang ZhangView author publicationsSearch author on:PubMed Google ScholarZhaochen ZhangView author publicationsSearch author on:PubMed Google ScholarZhonghua ZhangView author publicationsSearch author on:PubMed Google ScholarJian ZhangView author publicationsSearch author on:PubMed Google ScholarContributionsJ.Z. conceived the idea for the project and compiled the data. All authors were involved in collecting datasets. The first draft, the figures and tables were produced by X.W. All authors discussed and commented on the manuscript and contributed to the revised versions.Corresponding authorCorrespondence to
    Jian Zhang.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.Rights 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 articleWang, X., Chen, Y., Chen, Y. et al. Plant community data along elevational gradients in China’s 17 mountains.
    Sci Data (2025). https://doi.org/10.1038/s41597-025-06414-6Download citationReceived: 28 August 2025Accepted: 02 December 2025Published: 15 December 2025DOI: https://doi.org/10.1038/s41597-025-06414-6Share 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

    A global long-term daily multilayer soil moisture dataset derived from machine learning

    AbstractSoil moisture is a critical component of the Earth’s energy and water cycles. However, most existing products focus solely on surface layers, and continuous, high‐resolution datasets for deep soil horizons remain scarce. To address this gap, we generated a global, daily, seamless multilayer soil moisture dataset (SWSM) for the period 2002–2021 by leveraging a machine learning approach (XGBoost). The SWSM dataset provides estimates at a 0.05° spatial resolution for three depth horizons: 0–10 cm, 10–30 cm, and 30–60 cm. Rigorous validation against in situ observations demonstrated the dataset’s high accuracy, with Pearson correlation coefficients exceeding 0.90 and root mean square errors below 0.05 across all depths. A feature importance assessment verified the dataset’s physical consistency, revealing depth-dependent patterns aligned with established hydrological understanding. The SWSM dataset, with its long-term temporal coverage, fine spatial resolution, and multi-layer structure, is a valuable resource for applications in hydrologic modeling, agricultural water management, and climate change studies.

    Similar content being viewed by others

    Global soil moisture data derived through machine learning trained with in-situ measurements

    Article
    Open access
    12 July 2021

    High-resolution European daily soil moisture derived with machine learning (2003–2020)

    Article
    Open access
    14 November 2022

    Global long term daily 1 km surface soil moisture dataset with physics informed machine learning

    Article
    Open access
    17 February 2023

    Data availability

    The SWSM dataset generated in this study is openly available at Zenodo. Two repositories are provided: https://doi.org/10.5281/zenodo.15262116 and https://doi.org/10.5281/zenodo.15250534.
    Code availability

    The custom script used to read the NetCDF files in this study is publicly hosted on GitHub at the repository address: https://github.com/weizeyang1997/SWSM.
    ReferencesDorigo, W. et al. ESA CCI soil moisture for improved earth system understanding: state-of-the art and future directions. Remote Sens. Environ. 203, 185–215 (2017).
    Google Scholar 
    Yuan, Q., Xu, H., Li, T., Shen, H. & Zhang, L. Estimating surface soil moisture from satellite observations using a generalized regression neural network trained on sparse ground-based measurements in the continental U.S. Journal of Hydrology 580, 124351 (2020).
    Google Scholar 
    Dong, J., Akbar, R., Feldman, A. F., Gianotti, D. S. & Entekhabi, D. Land surfaces at the tipping‐point for water and energy balance coupling, https://doi.org/10.1029/2022WR032472.Zohaib, M., Kim, H. & Choi, M. Evaluating the patterns of spatiotemporal trends of root zone soil moisture in major climate regions in east Asia, https://doi.org/10.1002/2016JD026379.Shellito, P. J. et al. Assessing the impact of soil layer depth specification on the observability of modeled soil moisture and brightness temperature, https://doi.org/10.1175/JHM-D-19-0280.1 (2020).Song, P. et al. A 1&thinsp;km daily surface soil moisture dataset of enhanced coverage under all-weather conditions over china in 2003–2019. Earth Syst. Sci. Data 14, 2613–2637 (2022).
    Google Scholar 
    Zhang, N., Quiring, S. M. & Ford, T. W. Blending noah, SMOS, and in situ soil moisture using multiple weighting and sampling schemes, https://doi.org/10.1175/JHM-D-20-0119.1 (2021).Chen, Y., Feng, X. & Fu, B. An improved global remote-sensing-based surface soil moisture (RSSSM) dataset covering 2003–2018. Earth Syst. Sci. Data 13, 1–31 (2021).
    Google Scholar 
    Fisher, R. A. & Koven, C. D. Perspectives on the future of land surface models and the challenges of representing complex terrestrial systems. JAMES 12, e2018MS001453 (2020).
    Google Scholar 
    Tai, S.-L. et al. A 1&thinsp;km soil moisture dataset over eastern CONUS generated by assimilating SMAP data into the noah-MP land surface model. Earth Syst. Sci. Data 17, 4587–4611 (2025).
    Google Scholar 
    Feldman, A. F. et al. Remotely sensed soil moisture can capture dynamics relevant to plant water uptake, https://doi.org/10.1029/2022WR033814.Feldman, A. F. et al. Soil moisture profiles of ecosystem water use revealed with ECOSTRESS, https://doi.org/10.1029/2024GL108326.Liu, J., Rahmani, F., Lawson, K. & Shen, C. A multiscale deep learning model for soil moisture integrating satellite and In situ data, https://doi.org/10.1029/2021GL096847.Zhao, H., Montzka, C., Vereecken, H. & Franssen, H.-J. H. A comparative analysis of remote sensing soil moisture datasets fusion methods: novel LSTM approach versus widely used triple collocation technique. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 17, 16659–16671 (2024).
    Google Scholar 
    Hu, J., Deng, C., Zhang, Q. & Pang, A. Physics-informed neural networks enhanced by data augmentation: a novel framework for robust soil moisture estimation using multi-source data fusion. J. Hydrol. 663, 134320 (2025).
    Google Scholar 
    Chen, L. et al. Using remote sensing and machine learning to generate 100-cm soil moisture at 30-m resolution for the black soil region of China: implication for agricultural water management. Agric. Water Manage. 309, 109353 (2025).
    Google Scholar 
    Zhang, Y. et al. Generation of global 1&thinsp;km daily soil moisture product from 2000 to 2020 using ensemble learning. Earth Syst. Sci. Data 15, 2055–2079 (2023).
    Google Scholar 
    O, S. & Orth, R. Global soil moisture data derived through machine learning trained with in-situ measurements. Sci. Data 8, 170 (2021).
    Google Scholar 
    O, S., Orth, R., Weber, U. & Park, S. K. High-resolution european daily soil moisture derived with machine learning (2003–2020), https://doi.org/10.48550/arXiv.2205.10753 (2022).Han, Q. et al. Global long term daily 1 km surface soil moisture dataset with physics informed machine learning. Sci. Data 10, 101 (2023).
    Google Scholar 
    Padarian, J., McBratney, A. B. & Minasny, B. Game theory interpretation of digital soil mapping convolutional neural networks. Soil 6, 389–397 (2020).
    Google Scholar 
    Muñoz-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 
    Hersbach, H & Bell, B.: ERA5 hourly time-series data on single levels from 1940 to present. Copernicus Climate Change Service (C3S) Climate Data Store (CDS), https://doi.org/10.24381/cds.e2161bac (2025).Zhou, J., Liang, S., Cheng, J., Wang, Y. & Ma, J. The GLASS land surface temperature product. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 12, 493–507 (2019).
    Google Scholar 
    Ma, H. & Liang, S. Development of the GLASS 250-m leaf area index product (version 6) from MODIS data using the bidirectional LSTM deep learning model. Remote Sens. Environ. 273, 112985 (2022).
    Google Scholar 
    Friedl, M. & Sulla-Menashe, D. MODIS/terra+aqua land cover type yearly L3 global 0.05Deg CMG V061. NASA Land Processes Distributed Active Archive Center, https://doi.org/10.5067/MODIS/MCD12C1.061 (2022).Danielson, J. J. & Gesch, D. B. Global Multi-Resolution Terrain Elevation Data 2010 (GMTED2010). Open-File Report https://pubs.usgs.gov/publication/ofr20111073, https://doi.org/10.3133/ofr20111073 (2011).Hengl, T. et al. SoilGrids250m: global gridded soil information based on machine learning. PLOS One 12, e0169748 (2017).
    Google Scholar 
    Lehmann, P., Berli, M., Koonce, J. E. & Or, D. Surface evaporation in arid regions: insights from lysimeter decadal record and global application of a surface evaporation capacitor (SEC) model, https://doi.org/10.1029/2019GL083932.Beck, H. E. et al. Evaluation of 18 satellite- and model-based soil moisture products using in situ measurements from 826 sensors. Hydrol. Earth Syst. Sci. 25, 17–40 (2021).
    Google Scholar 
    Zhang, L. et al. Environmental factors driving evapotranspiration over a grassland in a transitional climate zone in China, https://doi.org/10.1002/met.2066.Yang, J., Li, Z., Zhai, P., Zhao, Y. & Gao, X. The influence of soil moisture and solar altitude on surface spectral albedo in arid area. Environ. Res. Lett. 15, 35010 (2020).
    Google Scholar 
    Hu, Y. et al. A physical method for downscaling land surface temperatures using surface energy balance theory. Remote Sens. Environ. 286, 113421 (2023).
    Google Scholar 
    Matsushima, D. Thermal inertia-based method for estimating soil moisture. in Soil Moisture, https://doi.org/10.5772/intechopen.80252 (IntechOpen, 2018).Zhang, J., Wang, W.-C. & Wu, L. Land‐atmosphere coupling and diurnal temperature range over the contiguous united states, https://doi.org/10.1029/2009GL037505.Lagos, L. O. et al. Surface energy balance model of transpiration from variable canopy cover and evaporation from residue-covered or bare soil systems: model evaluation. Irrig. Sci. 31, 135–150 (2013).
    Google Scholar 
    Alves, I. & do Rosário Cameira, M. Evapotranspiration estimation performance of root zone water quality model: evaluation and improvement. Agric. Water Manage. 57, 61–73 (2002).
    Google Scholar 
    Cisneros Vaca, C., van der Tol, C. & Ghimire, C. P. The influence of long-term changes in canopy structure on rainfall interception loss: a case study in speulderbos, the Netherlands. Hydrol. Earth Syst. Sci. 22, 3701–3719 (2018).
    Google Scholar 
    Hoek van Dijke, A. J. et al. Examining the link between vegetation leaf area and land–atmosphere exchange of water, energy, and carbon fluxes using FLUXNET data. Biogeosciences 17, 4443–4457 (2020).
    Google Scholar 
    Liu, Z. et al. Modeling the response of daily evapotranspiration and its components of a larch plantation to the variation of weather, soil moisture, and canopy leaf area index. J. Geophys. Res.: Atmos. 123, 7354–7374 (2018).
    Google Scholar 
    Chen, M., Willgoose, G. R. & Saco, P. M. Investigating the impact of leaf area index temporal variability on soil moisture predictions using remote sensing vegetation data. J. Hydrol. 522, 274–284 (2015).
    Google Scholar 
    Wang, Y., Yang, J., Chen, Y., Wang, A. & De Maeyer, P. The spatiotemporal response of soil moisture to precipitation and temperature changes in an arid region, china. Remote Sens. 10, 468 (2018).
    Google Scholar 
    Fan, L. et al. Mapping soil moisture at a high resolution over mountainous regions by integrating In situ measurements, topography data, and MODIS land surface temperatures. Remote Sens. 11, 656 (2019).
    Google Scholar 
    Lapides, D. A. et al. Inclusion of bedrock vadose zone in dynamic global vegetation models is key for simulating vegetation structure and function. Biogeosciences 21, 1801–1826 (2024).
    Google Scholar 
    Dai, Y. et al. A global high-resolution data set of soil hydraulic and thermal properties for land surface modeling. JAMES 11, 2996–3023 (2019).
    Google Scholar 
    Shangguan, W., Hengl, T., Mendes de Jesus, J., Yuan, H. & Dai, Y. Mapping the global depth to bedrock for land surface modeling. JAMES 9, 65–88 (2017).
    Google Scholar 
    Chen, L. & Dirmeyer, P. A. Impacts of land-use/land-cover change on afternoon precipitation over north america, https://doi.org/10.1175/JCLI-D-16-0589.1 (2017).Floriancic, M. G. et al. Potential for significant precipitation cycling by forest‐floor litter and deadwood, https://doi.org/10.1002/eco.2493.Li, Y. et al. Spatiotemporal impacts of land use land cover changes on hydrology from the mechanism perspective using SWAT model with time-varying parameters. Hydrol. Res. 50, 244–261 (2018).
    Google Scholar 
    Dorigo, W. et al. The international soil moisture network: serving earth system science for over a decade. Hydrol. Earth Syst. Sci. 25, 5749–5804 (2021).
    Google Scholar 
    Dorigo, W. A. et al. Global automated quality control of In situ soil moisture data from the international soil moisture network. Vadose Zone J. 12, vzj2012.97 (2013).
    Google Scholar 
    Li, Q. et al. A 1&thinsp;km daily soil moisture dataset over china using in situ measurement and machine learning. Earth Syst. Sci. Data 14, 5267–5286 (2022).
    Google Scholar 
    Chen, T. & Guestrin, C. XGBoost: a scalable tree boosting system. in Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining 785–794, https://doi.org/10.1145/2939672.2939785 (2016).Entekhabi, D., Reichle, R. H., Koster, R. D. & Crow, W. T. Performance Metrics for Soil Moisture Retrievals and Application Requirements. J. Hydrometeorol. 11, 832–840 (2010).
    Google Scholar 
    Wei, Z. High resolution daily multilayer soil moisture dataset 2002 to 2013 derived from integrated multi-source data fusion. Zenodo https://doi.org/10.5281/zenodo.15250534 (2025).Wei, Z. High resolution daily multilayer soil moisture dataset 2014 to 2021 derived from integrated multi-source data fusion. Zenodo https://doi.org/10.5281/zenodo.15262116 (2025).Beaudoing, H., Rodell, M. & Nasa/Gsfc/Hsl. GLDAS noah land surface model L4 3 hourly 0.25 x0.25 degree, version 2.1. NASA Goddard Earth Sciences Data and Information Services Center, https://doi.org/10.5067/E7TYRXPJKWOQ (2020).Miralles, D. G. et al. GLEAM4: global land evaporation and soil moisture dataset at 0.1° resolution from 1980 to near present. Sci. Data 12, 416 (2025).
    Google Scholar 
    Preimesberger, W., Stradiotti, P. & Dorigo, W. ESA CCI soil moisture GAPFILLED: an independent global gap-free satellite climate data record with uncertainty estimates. Earth Syst. Sci. Data 17, 4305–4329 (2025).
    Google Scholar 
    Download referencesAcknowledgementsThis work was supported by the National Natural Science Foundation of China (42271392), the Opening Foundation of Xi’an Key Laboratory of Territorial Spatial Information (3001023545016), and the Open Fund of the Key Laboratory of Natural Resources Monitoring and Supervision in Southern Hilly Region, Ministry of Natural Resources (NRMSSHR2022Y02, NRMSSHR2023Y03). We also gratefully acknowledge the QA4SM platform for providing critical soil moisture validation services; their efforts have greatly facilitated and enhanced the quality of our soil moisture research.Author informationAuthors and AffiliationsFaculty of Resources and Environmental Science, Hubei University, Wuhan, 430062, ChinaZeyang Wei, Lifei Wei, Qikai Lu & Shuang TianHubei Key Laboratory of Regional Development and Environmental Response, Hubei University, Wuhan, 430062, ChinaZeyang Wei, Lifei Wei, Qikai Lu & Shuang TianHubei Spatial Planning Research Institute, Wuhan, ChinaTing WangCollege of Geography and Environmental Science, Zhejiang Normal University, Jinhua, 321004, ChinaFei ZhangState Key Laboratory of Information Engineering in Surveying Mapping and Remote Sensing, Wuhan University, Wuhan, 430072, ChinaYanfei ZhongAuthorsZeyang WeiView author publicationsSearch author on:PubMed Google ScholarLifei WeiView author publicationsSearch author on:PubMed Google ScholarTing WangView author publicationsSearch author on:PubMed Google ScholarQikai LuView author publicationsSearch author on:PubMed Google ScholarShuang TianView author publicationsSearch author on:PubMed Google ScholarFei ZhangView author publicationsSearch author on:PubMed Google ScholarYanfei ZhongView author publicationsSearch author on:PubMed Google ScholarContributionsZ.W. and L.W. conceived the overall experiment; Z.W. conducted the entire experiment and wrote the manuscript; T.W. provided computational resources; Q.L. optimized the experimental procedure; S.T. assisted in data cleaning; F.Z. reviewed the manuscript; and Y.Z. guided the methodology and made revisions to the figures and manuscript. All authors have read and approved the final manuscript.Corresponding authorCorrespondence to
    Lifei Wei.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.Rights 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 articleWei, Z., Wei, L., Wang, T. et al. A global long-term daily multilayer soil moisture dataset derived from machine learning.
    Sci Data (2025). https://doi.org/10.1038/s41597-025-06436-0Download citationReceived: 25 April 2025Accepted: 10 December 2025Published: 15 December 2025DOI: https://doi.org/10.1038/s41597-025-06436-0Share 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 More

  • in

    Evolutionary adaptation of anaerobic and aerobic metabolism to high sulfide and hypoxic hydrothermal vent crab, Xenograpsus testudinatus

    AbstractThe vent crab, Xenograpsus testudinatus (xtcrab), is adapted to inhabit shallow-water, high sulfide and hypoxic hydrothermal vent. Our previous study revealed sulfide tolerance of vent xtcrabs which sulfide: quinone oxidoreductase (xtSQR) paralogs aid in sulfide detoxification. However, the mechanisms of how xtcrab adapts to high sulfide-hypoxic conditions in the vent area remain to be explored. In the present study, we tested the tolerance of xtcrab to sulfide-induced hypoxia, and investigated their aerobic and anaerobic responses in situ and in the laboratory. Comparisons were made to a non-vent, intertidal species, Thranita danae (tdcrab). We analyzed the several factors related to aerobic metabolism (SQR, cytochrome c [CYTC], complex IV [COXIV]), the product of anaerobic metabolism (hemolymph lactate levels) and glucose levels. Our results showed a higher survival tolerance to hypoxia of xtcrabs than tdcrabs. Hemolymph lactate levels increased more rapidly in xtcrabs than tdcrabs exposed to experimental hypoxia, revealing a rapid induction of anaerobic metabolism in hypoxic xtcrabs. Lactate measurement in xtcrabs returned from aquaria to original capture sites (vent habitats), further assessed the remarkable ability of xtcrabs to rapidly switch on and off their anaerobic metabolism. To assess aerobic metabolism, long-term exposure of xtcrabs to hydrothermal vent habitat increased gill xtCYTC transcripts and protein levels together with steadily enzymatic activity of COXIV. This revealed ability of xtcrabs to maintain functional capacity of aerobic respiration in hypoxia. Phylogenetic analysis showed that xtSQR paralogs in xtcrabs were more distant compared to tdSQR paralogs in tdcrabs. The increase of transcripts and enzymatic activity of gill xtSQR, and co-localization of xtSQR and xtCYTC also contribute to maintain aerobic metabolism by preventing sulfide toxicity on mitochondrial respiratory function. Overall, our study suggests that multiple strategies including detoxification of sulfide by gill xtSQR, and a quick/dynamic switch between aerobic and anaerobic metabolisms may play important roles in the metabolic adaptations of xtcrabs to extreme hydrothermal vent environment.

    Similar content being viewed by others

    Simultaneous aerobic and anaerobic respiration in hot spring chemolithotrophic bacteria

    Article
    Open access
    27 January 2025

    Structural basis for aerobic anoxygenic photosynthesis in the reaction center–light-harvesting 1 (RC–LH1) supercomplex of Dinoroseobacter shibae

    Article
    Open access
    14 November 2025

    When anaerobes encounter oxygen: mechanisms of oxygen toxicity, tolerance and defence

    Article

    28 June 2021

    Data availability

    The original data are available from Chi Chen and Ching-Fong Chang upon requests.
    ReferencesWannamaker, C. M. & Rice, J. A. Effects of hypoxia on movements and behavior of selected estuarine organisms from the southeastern united States. J. Exp. Mar. Biol. Ecol. 249 (2), 145–163. https://doi.org/10.1016/s0022-0981(00)00160-x (2000).
    Google Scholar 
    Giomi, F. & Beltramini, M. The molecular heterogeneity of hemocyanin: its role in the adaptive plasticity of crustacea. Gene 398 (1–2), 192–201. https://doi.org/10.1016/j.gene.2007.02.039 (2007).
    Google Scholar 
    McMahon, B. R. Respiratory and circulatory compensation to hypoxia in crustaceans. Respir Physiol. 128 (3), 349–364. https://doi.org/10.1016/S0034-5687(01)00311-5 (2001).
    Google Scholar 
    de Lima, T. M., Geihs, M. A., Nery, L. E. M. & Maciel, F. E. Air exposure behavior of the semiterrestrial crab Neohelice granulata allows tolerance to severe hypoxia but not prevent oxidative damage due to hypoxia–reoxygenation cycle. Physiol. Behav. 151, 97–101. https://doi.org/10.1016/j.physbeh.2015.07.013 (2015).
    Google Scholar 
    de Lima, T. M. et al. Emersion behavior of the semi-terrestrial crab Neohelice granulata during hypoxic conditions: Lactate as a trigger. Comp Biochem Physiol A Mol Integr Physiol, 252, 110835. (2021). https://doi.org/10.1016/j.cbpa.2020.110835 (2021).Hirota, S. et al. Structural basis of the lactate-dependent allosteric regulation of oxygen binding in arthropod Hemocyanin. J. Biol. Chem. 285 (25), 19338–19345 (2010).
    Google Scholar 
    Van Dover, C. L. The Ecology of deep-sea Hydrothermal Vents (Princeton University Press, 2000). https://doi.org/10.1515/9780691239477Powell, M. & Somero, G. Adaptations to sulfide by hydrothermal vent animals: sites and mechanisms of detoxification and metabolism. Biol. Bull. 171 (1), 274–290 (1986).
    Google Scholar 
    Chiu, L. et al. Shallow-water hydrothermal vent system as an extreme proxy for discovery of Microbiome significance in a crustacean holobiont. Front. Mar. Sci. 1670. https://doi.org/10.3389/fmars.2022.976255 (2022).Sun, Y. et al. Adaption to hydrogen sulfide-rich environments: strategies for active detoxification in deep-sea symbiotic mussels, Gigantidas platifrons. Sci. Total Environ. 804, 150054. https://doi.org/10.1016/j.scitotenv.2021.150054 (2022).
    Google Scholar 
    Mickel, T. J. & Childress, J. J. Effects of temperature, pressure, and oxygen concentration on the oxygen consumption rate of the hydrothermal vent crab bythograea thermydron (Brachyura). Physiol. Zool. 55 (2), 199–207 (1982).
    Google Scholar 
    Sanders, N. & Childress, J. Specific effects of thiosulphate and L-lactate on hemocyanin-O2 affinity in a brachyuran hydrothermal vent crab. Mar. Biol. 113 (2), 175–180. https://doi.org/10.1007/BF00347269 (1992).
    Google Scholar 
    Jeng, M. S., Ng, N. & Ng, P. Hydrothermal vent crabs feast on sea ‘snow’. Nature 432 (7020), 969–969. https://doi.org/10.1038/432969a (2004).
    Google Scholar 
    Chen, C. et al. Duplicated paralog of sulfide: Quinone oxidoreductase contributes to the adaptation to hydrogen sulfide-rich environment in the hydrothermal vent crab, xenograpsus testudinatus. Sci. Total Environ. 890, 164257. https://doi.org/10.1016/j.scitotenv.2023.164257 (2023).
    Google Scholar 
    Corrigan, E., Chen, C. J., Wang, B. S., Dufour, S. & Chang, C. F. Robustness of gametogenesis in the scleractinian coral, Tubastraea aurea, in the shallow-water hydrothermal vent field off Kueishan Island, Northeastern Taiwan. Sci. Total Environ. 992, 179901. https://doi.org/10.1016/j.scitotenv.2025.179901 (2025).
    Google Scholar 
    Chan, B. K. K. et al. Community structure of macrobiota and environmental parameters in shallow water hydrothermal vents off Kueishan Island, Taiwan. PLoS One. 11 (2), e0148675. https://doi.org/10.1371/journal.pone.0148675 (2016).
    Google Scholar 
    Mei, K. et al. Transformation, fluxes and impacts of dissolved metals from shallow water hydrothermal vents on nearby ecosystem offshore of Kueishantao (NE Taiwan). Sustainability 14 (3), 1754. https://doi.org/10.3390/su14031754 (2022).
    Google Scholar 
    Wang, Y. G. et al. Copepods as indicators of different water masses during the Northeast monsoon prevailing period in the Northeast Taiwan. Biology 11 (9), 1357. https://doi.org/10.3390/biology11091357 (2022).
    Google Scholar 
    Chiu, L., Wang, M. C., Wei, C. L., Lin, T. H. & Tseng, Y. C. A two-year physicochemical and acoustic observation reveals Spatiotemporal effects of earthquake‐induced shallow‐water hydrothermal venting on the surrounding environments. Limnol. Oceanogr. Lett. 9 (4), 423–432. https://doi.org/10.1002/lol2.10412 (2024).
    Google Scholar 
    Davidson, A. M., Tseng, L. C., Wang, Y. G. & Hwang, J. S. Mortality of mesozooplankton in an acidified ocean: investigating the impact of shallow hydrothermal vents across multiple monsoonal periods. Mar. Pollut Bull. 205, 116547. https://doi.org/10.1016/j.marpolbul.2024.116547 (2024).
    Google Scholar 
    Huang, Y. H. & Shih, H. T. Diversity of the swimming crabs (Crustacea: brachyura: Portunidae) from Dongsha Island, with a new record from Taiwan. J. Taiwan. Mus. 76 (3&4), 37–102. https://doi.org/10.6532/JNTM.202312_76(3_4).04 (2023).
    Google Scholar 
    Jansen, S. et al. Functioning of intertidal flats inferred from Temporal and Spatial dynamics of O2, H2S and pH in their surface sediment. Ocean. Dyn. 59 (2), 317–332 (2009).
    Google Scholar 
    Nicholls, P. & Kim, J. K. Sulphide as an inhibitor and electron donor for the cytochrome c oxidase system. Can. J. Biochem. 60 (6), 613–623. https://doi.org/10.1139/o82-076 (1982).
    Google Scholar 
    Khan, A. et al. Effects of hydrogen sulfide exposure on lung mitochondrial respiratory chain enzymes in rats. Toxicol. Appl. Pharmacol. 103 (3), 482–490. https://doi.org/10.1016/0041-008x(90)90321-k (1990).
    Google Scholar 
    Searcy, D. G. Metabolic integration during the evolutionary origin of mitochondria. Cell. Res. 13 (4), 229–238. https://doi.org/10.1038/sj.cr.7290168 (2003).
    Google Scholar 
    Vitvitsky, V. et al. Cytochrome c reduction by H2S potentiates sulfide signaling. ACS Chem. Biol. 13 (8), 2300–2307. https://doi.org/10.1021/acschembio.8b00463 (2018).
    Google Scholar 
    Morrison, B. R. S. An investigation into the effects of the piscicide antimycin A on the fish and invertebrates of a Scottish stream. Aquac Res. 10 (3), 111–122. https://doi.org/10.1111/j.1365-2109.1979.tb00262.x (1979).
    Google Scholar 
    Han, Y. H., Kim, S. H., Kim, S. Z. & Park, W. H. Antimycin A as a mitochondrial electron transport inhibitor prevents the growth of human lung cancer A549 cells. Oncol. Rep. 20 (3), 689–693. https://doi.org/10.3892/or_00000061 (2008).
    Google Scholar 
    29. Shahak, Y & Hauska, G. Sulfide Oxidation from Cyanobacteria to Humans: Sulfide–Quinone Oxidoreductase (SQR) in Sulfur Metabolism in Phototrophic Organisms. Advances in Photosynthesis and Respiration (ed. Hell, R., Dahl, C., Knaff, D & Leustek, T.) vol 27. Springer, Dordrecht. https://doi.org/10.1007/978-1-4020-6863-8_16(2008).
    Google Scholar 
    Marcia, M., Ermler, U., Peng, G. & Michel, H. A new structure-based classification of sulfide: Quinone oxidoreductases. Proteins 78 (5), 1073–1083. https://doi.org/10.1002/prot.22665 (2010).
    Google Scholar 
    Hu, M. Y. et al. Strong ion regulatory abilities enable the crab xenograpsus testudinatus to inhabit highly acidified marine vent systems. Front. Physiol. 7, 14. https://doi.org/10.3389/fphys.2016.00014 (2016).
    Google Scholar 
    Diaz, R. J. & Rosenberg, R. Spreading dead zones and consequences for marine ecosystems. Science 321 (5891), 926–929. https://doi.org/10.1126/science.1156401 (2008).
    Google Scholar 
    Isensee, K. et al. The ocean is losing its breath. In Ocean and Climate Scientific Notes. Vol. 2. 20–32 (2016). (2016).Arp, A. J. & Childress, J. J. Functional characteristics of the blood of the deep-sea hydrothermal vent brachyuran crab. Science 214 (4520), 559–561. https://doi.org/10.1126/science.214.4520.559 (1981).
    Google Scholar 
    Fredricks, K. T., Hubert, T. D., Amberg, J. J., Cupp, A. R. & Dawson, V. K. Chemical controls for an integrated pest management program. N Am. J. Fish. Manag. 41 (2), 289–300. https://doi.org/10.1002/nafm.10339 (2021).
    Google Scholar 
    Ott, K. C. & Antimycin A brief review of it’s chemistry, environmental fate, and toxicology. Biochem. Et Biophys. Acta. 1185, 1–9 (1994).
    Google Scholar 
    Saari, G. N. Antimycin-A species sensitivity distribution: perspectives for non-indigenous fish control. Manag Biol Invasions. 14(3). https://doi.org/%2010.3391/mbi.14.3.09 (2023). (2023).Thorpe, K. E., Taylor, A. C. & Huntingford, F. A. How costly is fighting? Physiological effects of sustained exercise and fighting in swimming crabs, Necora puber (L.)(Brachyura, Portunidae). Anim. Behav. 50 (6), 1657–1666. https://doi.org/10.1016/0003-3472(95)80019-0 (1995).
    Google Scholar 
    Cota-Ruiz, K., Peregrino-Uriarte, A. B., Felix-Portillo, M., Martnez-Quintana, J. A. & Yepiz-Plascencia, G. Expression of Fructose 1, 6-bisphosphatase and phosphofructokinase is induced in hepatopancreas of the white shrimp Litopenaeus vannamei by hypoxia. Mar. Environ. Res. 106, 1–9. https://doi.org/10.1016/j.marenvres.2015.02.003 (2015).
    Google Scholar 
    Reyes-Ramos, C. A. et al. Phosphoenolpyruvate Carboxykinase cytosolic and mitochondrial isoforms are expressed and active during hypoxia in the white shrimp Litopenaeus vannamei. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 226, 1–9. https://doi.org/10.1016/j.cbpb.2018.08.001 (2018).
    Google Scholar 
    Bao, J., Li, X., Yu, H. & Jiang, H. Respiratory metabolism responses of Chinese mitten crab, eriocheir sinensis and Chinese grass shrimp, palaemonetes sinensis, subjected to environmental hypoxia stress. Front. Physiol. 9, 1559. https://doi.org/10.3389/fphys.2018.01559 (2018).
    Google Scholar 
    Opie, L. H. & Lopaschuk, G. D. Fuels: Aerobic and Anaerobic metabolism. Heart Physiology: from Cell To Circulation 4th edn, 306–354 (Lippincott, Williams and Wilkins, 2004).Hervant, F., Garin, D., Mathieu, J. & Freminet, A. Lactate metabolism and glucose turnover in the subterranean crustacean niphargus virei during post-hypoxic recovery. J. Exp. Biol. 202 (5), 579–592. https://doi.org/10.1242/jeb.202.5.579 (1999).
    Google Scholar 
    Kabil, O. & Banerjee, R. Redox biochemistry of hydrogen sulfide. J. Biol. Chem. 285 (29), 21903–21907. https://doi.org/10.1074/jbc.R110.128363 (2010).
    Google Scholar 
    Kelly, J. L. et al. Mechanisms underlying adaptation to life in Hydroten sulfide-rich environments. Mol. Biol. Evol. 33 (6), 1419–1434. https://doi.org/10.1093/molbev/msw020 (2016).
    Google Scholar 
    Henry, R. P., Lucu, C., Onken, H. & Weihrauch, D. Multiple functions of the crustacean gill: osmotic/ionic regulation, acid-base balance, ammonia excretion, and bioaccumulation of toxic metals. Front. Physiol. 3, 431. https://doi.org/10.3389/fphys.2012.00431 (2012).
    Google Scholar 
    Ogunbona, O. B. & Claypool, S. M. Emerging roles in the biogenesis of cytochrome c oxidase for members of the mitochondrial carrier family. Front. Cell. Dev. Biol. 7, 3. https://doi.org/10.3389/fcell.2019.00003 (2019).
    Google Scholar 
    Herzig, R. P., Scacco, S. & Scarpulla, R. C. Sequential serum-dependent activation of CREB and NRF-1 leads to enhanced mitochondrial respiration through the induction of cytochrome c. J. Biol. Chem. 275 (17), 13134–13141. https://doi.org/10.1074/jbc.275.17.13134 (2000).
    Google Scholar 
    Jimenez-Gutierrez, L. R., Uribe-Carvajal, S., Sanchez-Paz, A., Chimeo, C. & Muhlia-Almazan, A. The cytochrome c oxidase and its mitochondrial function in the whiteleg shrimp Litopenaeus vannamei during hypoxia. J. Bioenerg Biomembr. 46, 189–196. https://doi.org/10.1007/s10863-013-9537-5 (2014).
    Google Scholar 
    Pfenninger, M. et al. Parallel evolution of Cox genes in H2S-tolerant fish as key adaptation to a toxic environment. Nat. Commun. 5 (1), 3873. https://doi.org/10.1038/ncomms4873 (2014).
    Google Scholar 
    Mellado, M., de Ana, A. M., Moreno, M. C., Martı́nez-A, C. & Rodrıguez-Frade, J. M. A potential immune escape mechanism by melanoma cells through the activation of chemokine-induced T cell death. Curr. Biol. 11 (9), 691–696. https://doi.org/10.1016/s0960-9822(01)00199-3 (2001).
    Google Scholar 
    Shi, Y. et al. Effects of salinity on survival, growth, haemolymph osmolality, gill Na+-K+‐ATPase activity, respiration and excretion of the sword Prawn Parapenaeopsis hardwickii. Aquac Res. 53 (2), 603–611. https://doi.org/10.1111/are.15604 (2022).
    Google Scholar 
    Kumar, S., Stecher, G. & Tamura, K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33 (7), 1870–1874. https://doi.org/10.1093/molbev/msw054 (2016).
    Google Scholar 
    Livak, K. J. & Schmittgen, T. D. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta delta C(T)) method. Methods 25 (4), 402–408. https://doi.org/10.1006/meth.2001.1262 (2001).
    Google Scholar 
    Download referencesAcknowledgementsThis work was supported by the Center of Excellence for the Oceans, NTOU from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE), Taiwan, the Yushan Scholar Program (Sylvie Dufour), MOE, Taiwan (MOE-113-YSFAG-0012-001-P2), and the National Science and Technology Council (NSTC 112-2313-B-019-008; 113-2313-B-019-014). We thank to captain Dai-Shiu Lan, SCUBA diving coach Jen-Wei Lu, and Jen-sheng Lu for xtcrabs collection. We thank the staff of Yung-Che Tseng’s laboratory at Academia Sinica for xtcrabs collection. We thank Ying-Syuan Lyu of Ching-Fong Chang’s laboratory at NTOU for xtcrabs collection. We thank Emily Corrigan for the English correction. Thanks to Jie-Lin Guo of Ching-Fong Chang’s laboratory at NTOU for tdcrabs collection.FundingThis work was supported by the Center of Excellence for the Oceans, NTOU from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE), Taiwan, the Yushan Scholar Program (Sylvie Dufour), MOE, Taiwan (MOE-113-YSFAG-0012-001-P2), and the National Science and Technology Council (NSTC 112-2313-B-019-008; 113-2313-B-019-014).Author informationAuthors and AffiliationsDepartment of Aquaculture, National Taiwan Ocean University, Keelung, TaiwanChi Chen, Guan-Chung Wu & Ching-Fong ChangCenter of Excellence for the Oceans, National Taiwan Ocean University, Keelung, TaiwanChi Chen, Guan-Chung Wu, Sylvie Dufour & Ching-Fong ChangInstitute of Cellular and Organismic Biology, Academia Sinica, Taipei, TaiwanYung-Che TsengBiology of Aquatic Organisms and Ecosystems (BOREA), Muséum National d’Histoire Naturelle, Sorbonne Université, CNRS, IRD, Paris, FranceSylvie DufourAuthorsChi ChenView author publicationsSearch author on:PubMed Google ScholarGuan-Chung WuView author publicationsSearch author on:PubMed Google ScholarYung-Che TsengView author publicationsSearch author on:PubMed Google ScholarSylvie DufourView author publicationsSearch author on:PubMed Google ScholarChing-Fong ChangView author publicationsSearch author on:PubMed Google ScholarContributionsChi Chen: conducted the sample collection, developed the methodologies in *xt* crab and *td* crab performed the experiments, data curation and analyses of data, and wrote the original draft. Guan-Chung Wu, Yung-Che Tseng and Ching-Fong Chang: developed the concept of the study, guided the experiments, and evaluated the data. Ching-Fong Chang: acquired funding, wrote, reviewed, and edited the paper. Sylvie Dufour: gave important input into conceptual and mechanistic insights, reviewed and edited the paper. All authors approved the paper.Corresponding authorsCorrespondence to
    Chi Chen, Guan-Chung Wu or Ching-Fong Chang.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Additional informationChing-Fong Chang: Lead contactPublisher’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 articleChen, C., Wu, GC., Tseng, YC. et al. Evolutionary adaptation of anaerobic and aerobic metabolism to high sulfide and hypoxic hydrothermal vent crab, Xenograpsus testudinatus.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-31968-1Download citationReceived: 04 September 2025Accepted: 05 December 2025Published: 15 December 2025DOI: https://doi.org/10.1038/s41598-025-31968-1Share 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
    KeywordsAerobic respiration; anaerobic respirationHydrogen sulfideCellular energyExtreme environmentAdaptation More

  • in

    Integrating microbial siderophores into concepts of plant iron nutrition

    AbstractIron is a crucial micronutrient for plants, but its availability in soil is often limited. Iron deficiency compromises plant growth, and low iron content in crops contributes substantially to the ‘hidden hunger’ that affects human health globally. The elucidation of Strategy I (reduction-based) and Strategy II (phytosiderophore-based) for iron acquisition was a milestone in plant biology and enabled the development of biofortification concepts. However, recent genetic evidence reveals that the boundary between the two strategies is blurred, with many plants possessing elements of both. Here we show that plant iron uptake mechanisms are more complex and diverse than the classical dichotomy suggests. We review evidence for this integrative view and highlight the critical role of microbial siderophores. We explain how plants access iron from microbial siderophores not only indirectly through Strategy I and II pathways but also via the direct uptake of iron–siderophore complexes, an overlooked mechanism that we introduce as Strategy III. We propose three potential routes for this direct uptake and conclude that harnessing Strategy III holds great potential for novel agricultural interventions to enhance iron biofortification and improve human health.

    Access through your institution

    Buy or subscribe

    This is a preview of subscription content, access via your institution

    Access options

    Access through your institution

    Access Nature and 54 other Nature Portfolio journals

    Get Nature+, our best-value online-access subscription

    $32.99 / 30 days

    cancel any time

    Learn more

    Subscribe to this journal

    Receive 12 digital issues and online access to articles

    $119.00 per year
    only $9.92 per issue

    Learn more

    Buy this articlePurchase on SpringerLinkInstant access to the full article PDF.USD 39.95Prices may be subject to local taxes which are calculated during checkout

    Additional access options:

    Log in

    Learn about institutional subscriptions

    Read our FAQs

    Contact customer support

    Fig. 1: Milestones in the discovery of molecular and physiological mechanisms underlying plant iron absorption.Fig. 2: Strategy I and II iron uptake mechanisms in plants and their partial co-occurrence in select species.Fig. 3: Phylogenetic distribution and functional traits of plant iron-promoting microorganisms.Fig. 4: Proposed mechanisms of how plants can use microbial siderophores for iron acquisition, including the emerging Strategy III.

    Similar content being viewed by others

    Biofortification’s contribution to mitigating micronutrient deficiencies

    Article

    02 January 2024

    Spatial IMA1 regulation restricts root iron acquisition on MAMP perception

    Article

    10 January 2024

    Functional mutants of Azospirillum brasilense elicit beneficial physiological and metabolic responses in Zea mays contributing to increased host iron assimilation

    Article
    Open access
    06 January 2021

    ReferencesAndrews, S. C., Robinson, A. K. & Rodríguez-Quiñones, F. Bacterial iron homeostasis. FEMS Microbiol. Rev. 27, 215–237 (2003).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ponka, P., Tenenbein, M. & Eaton, J. W. in Handbook on the Toxicology of Metals (eds Nordberg G. F. et al.) 879–902 (Elsevier, 2015).Lauderdale, J. M., Braakman, R., Forget, G., Dutkiewicz, S. & Follows, M. J. Microbial feedbacks optimize ocean iron availability. Proc. Natl Acad. Sci. USA 117, 4842–4849 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Colombo, C., Palumbo, G., He, J.-Z., Pinton, R. & Cesco, S. Review on iron availability in soil: interaction of Fe minerals, plants, and microbes. J. Soils Sediments 14, 538–548 (2014).Article 
    CAS 

    Google Scholar 
    Zuo, Y. & Zhang, F. Soil and crop management strategies to prevent iron deficiency in crops. Plant Soil 339, 83–95 (2011).Article 
    CAS 

    Google Scholar 
    Vélez-Bermúdez, I. C. & Schmidt, W. Plant strategies to mine iron from alkaline substrates. Plant Soil 483, 1–25 (2023).Article 

    Google Scholar 
    Cronin, S. J. F., Woolf, C. J., Weiss, G. & Penninger, J. M. The role of iron regulation in immunometabolism and immune-related disease. Front. Mol. Biosci. 6, 116 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Weffort, V. R. S. & Lamounier, J. A. Hidden hunger—a narrative review. J. Pediatr. (Rio J.) 100, S10–S17 (2024).Article 
    PubMed 

    Google Scholar 
    Jurkevitch, E. et al. Exploiting micronutrient interaction to optimize biofortification programs: the case for inclusion of selenium and iodine in the HarvestPlus program. Nutr. Rev. 62, 247–252 (2004).Article 

    Google Scholar 
    Röhmeld, V. & Marschner, H. Evidence for a specific uptake system for iron phytosiderophores in roots of grasses. Plant Physiol. 80, 175–180 (1986).Article 

    Google Scholar 
    Santi, S. & Schmidt, W. Dissecting iron deficiency-induced proton extrusion in Arabidopsis roots. N. Phytol. 183, 1072–1084 (2009).Article 
    CAS 

    Google Scholar 
    Robinson, N. J., Procter, C. M., Connolly, E. L. & Guerinot, M. L. A ferric-chelate reductase for iron uptake from soils. Nature 397, 694–697 (1999).Article 
    CAS 
    PubMed 

    Google Scholar 
    Vert, G. et al. IRT1, an Arabidopsis transporter essential for iron uptake from the soil and for plant growth. Plant Cell 14, 1223–1233 (2002).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Curie, C. et al. Maize yellow stripe1 encodes a membrane protein directly involved in Fe(III) uptake. Nature 409, 346–349 (2001).Article 
    CAS 
    PubMed 

    Google Scholar 
    Bashir, K. et al. Cloning and characterization of deoxymugineic acid synthase genes from graminaceous plants. J. Biol. Chem. 281, 32395–32402 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Nozoye, T. et al. Phytosiderophore efflux transporters are crucial for iron acquisition in graminaceous plants. J. Biol. Chem. 286, 5446–5454 (2011).Article 
    CAS 
    PubMed 

    Google Scholar 
    Durrett, T. P., Gassmann, W. & Rogers, E. E. The FRD3-mediated efflux of citrate into the root vasculature is necessary for efficient iron translocation. Plant Physiol. 144, 197–205 (2007).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yokosho, K., Yamaji, N., Ueno, D., Mitani, N. & Ma, J. F. OsFRDL1 is a citrate transporter required for efficient translocation of iron in rice. Plant Physiol. 149, 297–305 (2009).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kim, S. A. et al. Localization of iron in Arabidopsis seed requires the vacuolar membrane transporter VIT1. Science 314, 1295–1298 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Bashir, K. et al. The rice mitochondrial iron transporter is essential for plant growth. Nat. Commun. 2, 322 (2011).Article 
    PubMed 

    Google Scholar 
    Ling, H. Q., Bauer, P., Bereczky, Z., Keller, B. & Ganal, M. The tomato fer gene encoding a bHLH protein controls iron-uptake responses in roots. Proc. Natl Acad. Sci. USA 99, 13938–13943 (2002).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Colangelo, E. P. & Guerinot, M. L. The essential basic helix–loop–helix protein FIT1 is required for the iron deficiency response. Plant Cell 16, 3400–3412 (2004).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yuan, Y. et al. FIT interacts with AtbHLH38 and AtbHLH39 in regulating iron uptake gene expression for iron homeostasis in Arabidopsis. Cell Res. 18, 385–397 (2008).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ogo, Y. et al. The rice bHLH protein OsIRO2 is an essential regulator of the genes involved in Fe uptake under Fe-deficient conditions. Plant J. 51, 366–377 (2007).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wang, S. et al. A transcription factor OsbHLH156 regulates Strategy II iron acquisition through localising IRO2 to the nucleus in rice. N. Phytol. 225, 1247–1260 (2020).Article 
    CAS 

    Google Scholar 
    Li, X., Zhang, H., Ai, Q., Liang, G. & Yu, D. Two bHLH transcription factors, bHLH34 and bHLH104, regulate iron homeostasis in Arabidopsis thaliana. Plant Physiol. 170, 2478–2493 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jakoby, M., Wang, H.-Y., Reidt, W., Weisshaar, B. & Bauer, P. FRU (BHLH029) is required for induction of iron mobilization genes in Arabidopsis thaliana. FEBS Lett. 577, 528–534 (2004).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhang, H., Li, Y., Yao, X., Liang, G. & Yu, D. POSITIVE REGULATOR OF IRON HOMEOSTASIS1, OsPRI1, facilitates iron homeostasis. Plant Physiol. 175, 543–554 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kobayashi, T. et al. The transcription factor IDEF1 regulates the response to and tolerance of iron deficiency in plants. Proc. Natl Acad. Sci. USA 104, 19150–19155 (2007).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Grillet, L., Lan, P., Li, W., Mokkapati, G. & Schmidt, W. IRON MAN is a ubiquitous family of peptides that control iron transport in plants. Nat. Plants 4, 953–963 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Selote, D., Samira, R., Matthiadis, A., Gillikin, J. W. & Long, T. A. Iron-binding E3 ligase mediates iron response in plants by targeting Basic Helix–Loop–Helix transcription factors. Plant Physiol. 167, 273–286 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Salahudeen, A. A. et al. An E3 ligase possessing an iron-responsive hemerythrin domain is a regulator of iron homeostasis. Science 326, 722–726 (2009).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vashisht, A. A. et al. Control of iron homeostasis by an iron-regulated ubiquitin ligase. Science 326, 718–721 (2009).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Long, T. et al. The bHLH transcription factor POPEYE regulates response to iron deficiency in Arabidopsis roots. Plant Cell 22, 2219–2236 (2010).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stacey, M. G. et al. The Arabidopsis AtOPT3 protein functions in metal homeostasis and movement of iron to developing seeds. Plant Physiol. 146, 323–324 (2008).Article 

    Google Scholar 
    Li, Y. et al. IRON MAN interacts with BRUTUS to maintain iron homeostasis in Arabidopsis. Proc. Natl Acad. Sci. USA 118, e2109063118 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kumar, R. K. et al. Iron-nicotianamine transporters are required for proper long distance iron signaling. Plant Physiol. 175, 1254–1268 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Brown, J. C. & Jolley, V. D. Strategy I and strategy II mechanisms affecting iron availability to plants may be established too narrow or limited. J. Plant Nutr. 11, 1077–1098 (1988).Article 
    CAS 

    Google Scholar 
    Martín-Barranco, A., Thomine, S., Vert, G. & Zelazny, E. A quick journey into the diversity of iron uptake strategies in photosynthetic organisms. Plant Signal. Behav. 16, 1975088 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chao, Z. & Chao, D. Similarities and differences in iron homeostasis strategies between graminaceous and nongraminaceous plants. N. Phytol. 236, 1655–1660 (2022).Article 
    CAS 

    Google Scholar 
    Ishimaru, Y. et al. Rice plants take up iron as an Fe 3+ -phytosiderophore and as Fe 2+. Plant J. 45, 335–346 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zuo, Y. M., Zhang, F. S., Li, X. L. & Cao, Y. P. Studies on the improvement in iron nutrition of peanut by intercropping with maize on a calcareous soil. Plant Soil 220, 13–25 (2000).Article 
    CAS 

    Google Scholar 
    Zuo, Y., Li, X., Cao, Y., Zhang, F. & Christie, P. Iron nutrition of peanut enhanced by mixed cropping with maize: possible role of root morphology and rhizosphere microflora. J. Plant Nutr. 26, 2093–2110 (2003).Article 
    CAS 

    Google Scholar 
    Guo, X. et al. Dynamics in the rhizosphere and iron-uptake gene expression in peanut induced by intercropping with maize: role in improving iron nutrition in peanut. Plant Physiol. Biochem. 76, 36–43 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Dai, J. et al. From Leguminosae/Gramineae intercropping systems to see benefits of intercropping on iron nutrition. Front. Plant Sci. 10, 605 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Xiong, H. et al. Molecular evidence for phytosiderophore-induced improvement of iron nutrition of peanut intercropped with maize in calcareous soil. Plant Cell Environ. 38, 1888–1902 (2013).Article 

    Google Scholar 
    He, R. et al. SIDERITE: unveiling hidden siderophore diversity in the chemical space through digital exploration. iMeta 3, e192 (2024).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hider, R. C. & Kong, X. Chemistry and biology of siderophores. Nat. Prod. Rep. 27, 637–657 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kramer, J., Özkaya, Ö & Kümmerli, R. Bacterial siderophores in community and host interactions. Nat. Rev. Microbiol. 18, 152–163 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Masalha, J., Kosegarten, H., Elmaci, O. & Mengel, K. The central role of microbial activity for iron acquisition in maize and sunflower. Biol. Fertil. Soils 30, 433–439 (2000).Article 
    CAS 

    Google Scholar 
    Rroço, E., Kosegarten, H., Harizaj, F., Imani, J. & Mengel, K. The importance of soil microbial activity for the supply of iron to sorghum and rape. Eur. J. Agron. 19, 487–493 (2003).Article 

    Google Scholar 
    Jin, C. W., He, Y. F., Tang, C. X., Wu, P. & Zheng, S. J. Mechanisms of microbially enhanced Fe acquisition in red clover (Trifolium pratense L.). Plant Cell Environ. 29, 888–897 (2006).Article 
    PubMed 

    Google Scholar 
    Wang, N. et al. Microbiome convergence enables siderophore-secreting-rhizobacteria to improve iron nutrition and yield of peanut intercropped with maize. Nat. Commun. 15, 839 (2024).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gu, S. et al. Competition for iron drives phytopathogen control by natural rhizosphere microbiomes. Nat. Microbiol. 5, 1002–1010 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Singh, D. et al. Prospecting endophytes from different Fe or Zn accumulating wheat genotypes for their influence as inoculants on plant growth, yield, and micronutrient content. Ann. Microbiol. 68, 815–833 (2018).Article 
    CAS 

    Google Scholar 
    Vansuyt, G., Robin, A., Briat, J. F., Curie, C. & Lemanceau, P. Iron acquisition from Fe-pyoverdine by Arabidopsis thaliana. Mol. Plant Microbe Interact. 20, 441–447 (2007).Article 
    CAS 
    PubMed 

    Google Scholar 
    Trapet, P. et al. The Pseudomonas fluorescens siderophore pyoverdine weakens Arabidopsis thaliana defense in favor of growth in iron-deficient conditions. Plant Physiol. 171, 675–693 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Avoscan, L. et al. Iron status and root cell morphology of Arabidopsis thaliana as modified by a bacterial ferri-siderophore. Physiol. Plant. 176, e14223 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Shirley, M., Avoscan, L., Bernaud, E., Vansuyt, G. & Lemanceau, P. Comparison of iron acquisition from Fe–pyoverdine by strategy I and strategy II plants. Botany 89, 731–735 (2011).Article 
    CAS 

    Google Scholar 
    Omidvari, M., Sharifi, R. A., Ahmadzadeh, M. & Dahaji, P. A. Role of fluorescent pseudomonads siderophore to increase bean growth factors. J. Agric. Sci. 2, 242–247 (2010).
    Google Scholar 
    Braun, V. & Killmann, H. Bacterial solutions to the iron-supply problem. Trends Biochem. Sci. 24, 104–109 (1999).Article 
    CAS 
    PubMed 

    Google Scholar 
    Robin, A. et al. in Advances in Agronomy, Vol. 99 (ed. Sparks D. L.) 183–225 (Elsevier, 2008).Crowley, D. E., Wang, Y. C., Reid, C. P. P. & Szaniszlo, P. J. Mechanisms of iron acquisition from siderophores by microorganisms and plants. Front. Microbiol. 130, 179–198 (1991).CAS 

    Google Scholar 
    Rai, V., Fisher, N., Duckworth, O. W. & Baars, O. Extraction and detection of structurally diverse siderophores in soil. Front. Microbiol. 11, 581508 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Harbort, C. J. et al. Root-secreted coumarins and the microbiota interact to improve iron nutrition in Arabidopsis. Cell Host Microbe 28, 825–837 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Johnson, G. V., Lopez, A. & Foster, N. L. Reduction and transport of Fe from siderophores—reduction of siderophores and chelates and uptake and transport of iron by cucumber seedlings. Plant Soil 241, 27–33 (2002).Article 
    CAS 

    Google Scholar 
    Yehuda, Z., Shenker, M., Hadar, Y. & Chen, Y. Remedy of chlorosis induced by iron deficiency in plants with the fungal siderophore rhizoferrin. J. Plant Nutr. 23, 1991–2006 (2000).Article 
    CAS 

    Google Scholar 
    Bienfait, H. F. Prevention of stress in iron metabolism of plants. Acta Bot. Neerl. 38, 105–129 (1989).Article 
    CAS 

    Google Scholar 
    Boukhalfa, H. & Crumbliss, A. Chemical aspects of siderophore mediated iron transport. Biometals 15, 325–339 (2002).Article 
    CAS 
    PubMed 

    Google Scholar 
    Jin, C. W., Ye, Y. Q. & Zheng, S. J. An underground tale: contribution of microbial activity to plant iron acquisition via ecological processes. Ann. Bot. 113, 7–18 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yehuda, Z. et al. The role of ligand exchange in the uptake of iron from microbial siderophores by gramineous plants. Plant Physiol. 112, 1273–1280 (1996).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jurkevitch, E., Hadar, Y., Chen, Y., Chino, M. & Mori, S. Indirect utilization of the phytosiderophore mugineic acid as an iron source to rhizosphere fluorescent Pseudomonas. Biometals 6, 119–123 (1993).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ahmed, E. & Holmström, S. J. M. Siderophores in environmental research: roles and applications. Microb. Biotechnol. 7, 196–208 (2014).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nishizawa, N. & Mori, S. Invagination of plasmalemma: its role in the absorption of macromolecules in rice roots. Plant Cell Physiol. 18, 767–782 (1977).
    Google Scholar 
    Mori, S. Iron acquisition by plants. Curr. Opin. Plant Biol. 2, 250–253 (1999).Article 
    CAS 
    PubMed 

    Google Scholar 
    Singh, P., Kumar, R., Khan, A., Singh, A. & Srivastava, A. Bacillibactin siderophore induces iron mobilisation responses inside aerobic rice variety through YSL15 transporter. Rhizosphere 27, 100724 (2023).Article 

    Google Scholar 
    Singh, P. et al. In silico analysis of comparative affinity of phytosiderophore and bacillibactin for iron uptake by YSL15 and YSL18 receptors of Oryza sativa. J. Biomol. Struct. Dyn. 41, 2733–2746 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Murata, Y. et al. A specific transporter for iron(III)–phytosiderophore in barley roots. Plant J. 46, 563–572 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Chen, L. M., Dick, W. A. & Streeter, J. G. Production of aerobactin by microorganisms from a compost enrichment culture and soybean utilization. J. Plant Nutr. 23, 2047–2060 (2000).Article 
    CAS 

    Google Scholar 
    Dahhan, D. A. & Bednarek, S. Y. Advances in structural, spatial, and temporal mechanics of plant endocytosis. FEBS Lett. 596, 2269–2287 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hayat, R. et al. Endocytosis-mediated siderophore uptake as a strategy for Fe acquisition in diatoms. Sci. Adv. 4, eaar4536 (2018).Article 

    Google Scholar 
    Diggle, S. P. et al. The Pseudomonas aeruginosa 4-quinolone signal molecules HHQ and PQS play multifunctional roles in quorum sensing and iron entrapment. Chem. Biol. 14, 87–96 (2007).Article 
    CAS 
    PubMed 

    Google Scholar 
    Lin, J. et al. A Pseudomonas T6SS effector recruits PQS-containing outer membrane vesicles for iron acquisition. Nat. Commun. 8, 14888 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chaney, R. L., Brown, J. C. & Tiffin, L. O. Obligatory reduction of ferric chelates in iron uptake by soybeans. Plant Physiol. 50, 208–213 (1972).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Eide, D., Broderius, M., Fett, J. & Guerinot, M. L. A novel iron-regulated metal transporter from plants identified by functional expression in yeast. Proc. Natl Acad. Sci. USA 93, 5624–5628 (1996).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rodríguez-Celma, J. et al. Root responses of Medicago truncatula plants grown in two different iron deficiency conditions: changes in root protein profile and riboflavin biosynthesis. J. Proteome Res. 10, 2590–2601 (2011).Article 
    PubMed 

    Google Scholar 
    Rodriguez-Celma, J. et al. Mutually exclusive alterations in secondary metabolism are critical for the uptake of insoluble iron compounds by Arabidopsis and Medicago truncatula. Plant Physiol. 162, 1473–1485 (2013).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fourcroy, P. et al. Involvement of the ABCG37 transporter in secretion of scopoletin and derivatives by Arabidopsis roots in response to iron deficiency. N. Phytol. 201, 155–167 (2014).Article 
    CAS 

    Google Scholar 
    Schmid, N. B. et al. Feruloyl-CoA 6′-hydroxylase1-dependent coumarins mediate iron acquisition from alkaline substrates in Arabidopsis. Plant Physiol. 164, 160–172 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Schmidt, H. et al. Metabolome analysis of Arabidopsis thaliana roots identifies a key metabolic pathway for iron acquisition. PLoS ONE 9, e102444 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rajniak, J. et al. Biosynthesis of redox-active metabolites in response to iron deficiency in plants. Nat. Chem. Biol. 14, 442–450 (2018).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Robe, K. et al. Coumarin-facilitated iron transport: an IRT1-independent strategy for iron acquisition in Arabidopsis thaliana. Plant Commun. 6, 101431 (2025).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Takagi, S. Naturally occurring iron-chelating compounds in oat- and rice-root washings. J. Soil Sci. Plant Nutr. 22, 423–433 (1976).Article 
    CAS 

    Google Scholar 
    Takagi, S., Nomoto, K. & Takemoto, T. Physiological aspect of mugineic acid, a possible phytosiderophore of graminaceous plants. J. Plant Nutr. 7, 469–477 (1984).Article 
    CAS 

    Google Scholar 
    Mori, S. & Nishizawa, N. K. Methionine as a dominant precursor of phytosiderophores in Graminaceae plants. Plant Cell Physiol. 28, 1081–1092 (1987).CAS 

    Google Scholar 
    Shojima, S. et al. Biosynthesis of phytosiderophores: in vitro biosynthesis of 2′-deoxymugineic acid from L-methionine and nicotianamine. Plant Physiol. 93, 1497–1503 (1990).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Higuchi, K. et al. Cloning of nicotianamine synthase genes, novel genes involved in the biosynthesis of phytosiderophores. Plant Physiol. 119, 471–479 (1999).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Inoue, H. et al. Three rice nicotianamine synthase genes, OsNAS1, OsNAS2, and OsNAS3 are expressed in cells involved in long-distance transport of iron and differentially regulated by iron. Plant J. 36, 366–381 (2003).Article 
    CAS 
    PubMed 

    Google Scholar 
    Takahashi, M. et al. Cloning two genes for nicotianamine aminotransferase, a critical enzyme in iron acquisition (Strategy II) in graminaceous plants. Plant Physiol. 121, 947–956 (1999).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Inoue, H. et al. Identification and localisation of the rice nicotianamine aminotransferase gene OsNAAT1 expression suggests the site of phytosiderophore synthesis in rice. Plant Mol. Biol. 66, 193–203 (2008).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ma, J. F., Shinada, T., Matsuda, C. & Nomoto, K. Biosynthesis of phytosiderophores, mugineic acids, associated with methionine cycling. J. Biol. Chem. 270, 16549–16554 (1995).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ma, J. F. & Nomoto, K. Two related biosynthetic pathways of mugineic acids in gramineous plants. Plant Physiol. 102, 373–378 (1993).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Francis, J., Madinaveitia, J., Macturk, H. M. & Snow, G. A. Isolation from acid-fast bacteria of a growth-factor for Mycobacterium johnei and of a precursor of phthiocol. Nature 163, 365–366 (1949).Article 
    CAS 
    PubMed 

    Google Scholar 
    Neilands, J. B. in Bioinorganic Chemistry—II, Vol. 261 (ed. Raymond, K. N.) 3–32 (American Chemical Society, 1977).Powell, P. E., Szaniszlo, P. J., Cline, G. R. & Reid, C. P. P. Hydroxamate siderophores in the iron nutrition of plants. J. Plant Nutr. 5, 653–673 (1982).Article 
    CAS 

    Google Scholar 
    Download referencesAcknowledgementsWe thank the National Natural Science Foundation of China (grant nos 42325704, 32372810, 42577142 and 32573128), the Disciplinary Breakthrough Project of Ministry of Education (MOE, #00975101), the National Key Research and Development Program of China (grant nos 2022YFD1901500/2022YFD1901501 and 2023YFD1700203), the Tianchi Talent Introduction Program of Xinjiang Autonomous Region, China (2023—‘2+5’), the Tingzhou Talent Introduction Program of Changji Autonomous Region, China (2023) and the Swiss National Science Foundation (grant no. 310030_212266) for funding. We thank S. J. Zheng from Zhejiang University and J. F. Ma from Okayama University for valuable discussions and suggestions.Author informationAuthor notesThese authors contributed equally: Shaohua Gu, Nanqi Wang.Authors and AffiliationsCollege of Resources and Environmental Sciences, State Key Laboratory of Nutrient Use and Management, National Academy of Agriculture Green Development, China Agricultural University, Beijing, ChinaShaohua Gu, Tianqi Wang, Fusuo Zhang & Yuanmei ZuoNational Citrus Engineering Research Center, Chongqing Key Laboratory of Citrus, Citrus Research Institute, Southwest University, Chongqing, ChinaNanqi WangJiangsu Provincial Key Lab for Organic Solid Waste Utilization, Key Lab of Organic-Based Fertilizers of China, Nanjing Agricultural University, Nanjing, ChinaYiran Zheng, Qirong Shen & Zhong WeiDepartment of Quantitative Biomedicine, University of Zurich, Zurich, SwitzerlandRolf KümmerliAuthorsShaohua GuView author publicationsSearch author on:PubMed Google ScholarNanqi WangView author publicationsSearch author on:PubMed Google ScholarYiran ZhengView author publicationsSearch author on:PubMed Google ScholarTianqi WangView author publicationsSearch author on:PubMed Google ScholarQirong ShenView author publicationsSearch author on:PubMed Google ScholarFusuo ZhangView author publicationsSearch author on:PubMed Google ScholarRolf KümmerliView author publicationsSearch author on:PubMed Google ScholarZhong WeiView author publicationsSearch author on:PubMed Google ScholarYuanmei ZuoView author publicationsSearch author on:PubMed Google ScholarContributionsY.Z. and Z.W. developed the concept. S.G., N.W., T.W. and Y.Z. performed the literature search and prepared the figures. F.Z. and Q.S. provided some intellectual input for this manuscript. S.G., N.W., R.K., Y.Z. and Z.W. wrote the manuscript with contributions and input from all authors.Corresponding authorsCorrespondence to
    Zhong Wei or Yuanmei Zuo.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Plants thanks Takanori Kobayashi and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

    Additional informationPublisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.Supplementary informationSupplementary InformationSupplementary Tables 1–4.Rights and permissionsSpringer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.Reprints and permissionsAbout this articleCite this articleGu, S., Wang, N., Zheng, Y. et al. Integrating microbial siderophores into concepts of plant iron nutrition.
    Nat. Plants (2025). https://doi.org/10.1038/s41477-025-02171-xDownload citationReceived: 04 July 2025Accepted: 06 November 2025Published: 15 December 2025Version of record: 15 December 2025DOI: https://doi.org/10.1038/s41477-025-02171-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

  • in

    Impact of mixing duration on growth and nutrient removal efficiency of Scenedesmus sp. in a novel raceway pond system

    AbstractRaceway ponds are regarded as a popular and cost-effective method for microalgae cultivation; however, their performance is strongly influenced by hydrodynamic conditions. Conventional paddlewheel driven systems are restricted to low operating velocities to avoid culture spilling out which often leads to reduced mixing and stagnant zone formation. In this study, a raceway pond was designed with the inclusion of curved slits at the bent zones and submersible pump as an alternative mixing device to prevent culture overflow, improve flow stability and minimize dead zones. This novel integration of structural modifications and pump-based mixing represents a significant advancement over traditional paddlewheel systems by providing higher velocities, enhanced circulation and more uniform algal growth conditions. Experiments were conducted to evaluate the effect of mixing durations in 6 L raceway ponds under identical environmental conditions. The raceway systems permitted a broader velocity range (0.10–0.45 m s−1) without spillage. The system with continuous 24 h mixing compared to 20 and 16 h mixing resulted in the highest biomass productivity of 1.01 g L−1 d−1 and maximum nutrient removal rates of 5.18 mg L−1 d-1 and 3.41 mg L−1 d−1 for NO3− and PO43−, respectively. Submersible-pump configured open raceway pond achieved comparable or higher biomass yield, lower energy consumption with a net energy efficiency of 62%, demonstrating its practicality and cost-effectiveness as a viable alternative to conventional paddlewheel driven systems for large-scale Scenedesmus sp. cultivation.

    Similar content being viewed by others

    Physical structure of the environment contributes to the development of diversity of microalgal assemblages

    Article
    Open access
    12 June 2024

    Comprehensive dataset of shotgun metagenomes from oxygen stratified freshwater lakes and ponds

    Article
    Open access
    14 May 2021

    A sandponics comparative study investigating different sand media based integrated aqua vegeculture systems using desalinated water

    Article
    Open access
    30 June 2022

    Data availability

    Data will be made available on request. Requests for data should be directed to Dr. Rashid Iftikhar ([email protected]; [email protected]).
    ReferencesVadiveloo, A. & Moheimani, N. Effect of continuous and daytime mixing on Nannochloropsis growth in raceway ponds. Algal Res. 33, 190–196. https://doi.org/10.1016/j.algal.2018.05.018 (2018).
    Google Scholar 
    Wijffels, R. H., Kruse, O. & Hellingwerf, K. J. Potential of industrial biotechnology with cyanobacteria and eukaryotic microalgae. Curr. Opin. Biotechnol. 24, 405–413. https://doi.org/10.1016/j.copbio.2013.04.004 (2013).
    Google Scholar 
    Vanthoor-Koopmans, M., Wijffels, R. H., Barbosa, M. J. & Eppink, M. H. M. Biorefinery of microalgae for food and fuel. Bioresour. Technol. 135, 142–149. https://doi.org/10.1016/j.biortech.2012.10.135 (2013).
    Google Scholar 
    Valverde, F., Romero-Campero, F. J., León, R., Guerrero, M. G. & Serrano, A. New challenges in microalgae biotechnology. Eur. J. Protistol. 55, 95–101. https://doi.org/10.1016/j.ejop.2016.03.002 (2016).
    Google Scholar 
    Borowitzka, M. Large-scale algal culture systems: the next generation. Australasian Biotechnology (1994).Jorquera, O., Kiperstok, A., Sales, E. A., Embiruçu, M. & Ghirardi, M. L. Comparative energy life-cycle analyses of microalgal biomass production in open ponds and photobioreactors. Bioresour. Technol. 101, 1406–1413. https://doi.org/10.1016/j.biortech.2009.09.038 (2010).
    Google Scholar 
    Kumar, K., Mishra, S. K., Shrivastav, A., Park, M. S. & Yang, J. W. Recent trends in the mass cultivation of algae in raceway ponds. Renew. Sustain. Energy Rev. 51, 875–885. https://doi.org/10.1016/j.rser.2015.06.033 (2015).
    Google Scholar 
    Borowitzka, M. A. & Moheimani, N. R. Open Pond Culture Systems. in Algae for Biofuels and Energy (eds Borowitzka, M. A. & Moheimani, N. R.) 133–152 (2013). https://doi.org/10.1007/978-94-007-5479-9_8Lundquist, T. J., Woertz, I. C., Quinn, N. W. T. & Benemann, J. R. A Realistic Technology and Engineering Assessment of Algae Biofuel Production. (2010).Rogers, J. N. et al. A critical analysis of paddlewheel-driven raceway ponds for algal biofuel production at commercial scales. Algal Res. 4, 76–88. https://doi.org/10.1016/j.algal.2013.11.007 (2014).
    Google Scholar 
    Stephens, E. et al. An economic and technical evaluation of microalgal biofuels. Nat. Biotechnol. 28, 126–128. https://doi.org/10.1038/nbt0210-126 (2010).
    Google Scholar 
    Kusmayadi, A., Philippidis, G. P. & Yen, H. W. Application of computational fluid dynamics to raceways combining paddlewheel and CO2 spargers to enhance microalgae growth. J. Biosci. Bioeng. 129, 93–98. https://doi.org/10.1016/j.jbiosc.2019.06.013 (2020).
    Google Scholar 
    Cheng, J., Yang, Z., Ye, Q., Zhou, J. & Cen, K. Enhanced flashing light effect with up-down chute baffles to improve microalgal growth in a raceway pond. Bioresour. Technol. 190, 29–35. https://doi.org/10.1016/j.biortech.2015.04.050 (2015).
    Google Scholar 
    Huang, J. et al. Design and optimization of a novel airlift-driven sloping raceway pond with numerical and practical experiments. Algal Res. 20, 164–171. https://doi.org/10.1016/j.algal.2016.09.023 (2016).
    Google Scholar 
    Jebali, A., Acién, F. G., Sayadi, S. & Molina-Grima, E. Utilization of centrate from urban wastewater plants for the production of scenedesmus sp. in a raceway-simulating reactor. J. Environ. Manage. 211, 112–124. https://doi.org/10.1016/j.jenvman.2018.01.043 (2018).
    Google Scholar 
    Chakraborty, B., Gayen, K. & Bhowmick, T. K. Transition from synthetic to alternative media for microalgae cultivation: A critical review. Sci. Total Environ. 897, 165412. https://doi.org/10.1016/j.scitotenv.2023.165412 (2023).
    Google Scholar 
    Saleem, S. et al. Operation of microalgal horizontal twin layer system for treatment of real wastewater and production of lipids. J. Water Process. Eng. 48, 102932. https://doi.org/10.1016/j.jwpe.2022.102932 (2022).
    Google Scholar 
    Bulynina, S. S., Ziganshina, E. E. & Ziganshin, A. M. Growth Efficiency of Chlorella sorokiniana in Synthetic Media and Unsterilized Domestic Wastewater. BioTech 12, 53 (2023). https://doi.org/10.3390/biotech12030053Malek, A., Zullo, L. C. & Daoutidis, P. Modeling and dynamic optimization of microalgae cultivation in outdoor open ponds. Ind. Eng. Chem. Res. 55, 3327–3337. https://doi.org/10.1021/acs.iecr.5b03209 (2016).
    Google Scholar 
    Matanguihan, A. E. D. et al. Design, fabrication, and performance evaluation of open raceway ponds for the cultivation of chlorella vulgaris Beijerinck in the Philippines. Philippine J. Sci. 149, 353–362. https://doi.org/10.56899/149.02.13 (2020).
    Google Scholar 
    Ahmad, S., Pathak, V. V., Kothari, R., Kumar, A. & Naidu Krishna, S. B. Optimization of nutrient stress using C. pyrenoidosa for lipid and biodiesel production in integration with remediation in dairy industry wastewater using response surface methodology. 3 Biotech. 8, 326. https://doi.org/10.1007/s13205-018-1342-8 (2018).
    Google Scholar 
    Xu, C., Wang, L., Liu, Z., Cai, G. & Zhan, J. Nitrogen and phosphorus removal efficiency and algae viability in an immobilized algae and bacteria symbiosis system with Pink luminescent filler. Water Sci. Technol. 85, 104–115. https://doi.org/10.2166/wst.2021.606 (2022).
    Google Scholar 
    Liu, Y. et al. Treatment of real aquaculture wastewater from a fishery utilizing phytoremediation with microalgae. J. Chem. Technol. Biotechnol. 94, 900–910. https://doi.org/10.1002/jctb.5837 (2019).
    Google Scholar 
    Saleem, S., Sheikh, Z., Iftikhar, R. & Zafar, M. I. Eco-friendly cultivation of microalgae using a horizontal twin layer system for treatment of real solid waste leachate. J. Environ. Manage. 351, 119847. https://doi.org/10.1016/j.jenvman.2023.119847 (2024).
    Google Scholar 
    APHA. Standard Methods for the Examination of Water and Wastewater. (2017).Hemalatha, M., Sravan, J. S., Min, B. & Venkata Mohan, S. Microalgae-biorefinery with cascading resource recovery design associated to dairy wastewater treatment. Bioresour. Technol. 284, 424–429. https://doi.org/10.1016/j.biortech.2019.03.106 (2019).
    Google Scholar 
    Richmond, A. Biological Principles of Mass Cultivation of Photoautotrophic Microalgae. in Handbook of Microalgal Culture 169–204 (2013). https://doi.org/10.1002/9781118567166.ch11Hadiyanto, H., Elmore, S., Van Gerven, T. & Stankiewicz, A. Hydrodynamic evaluations in high rate algae pond (HRAP) design. Chem. Eng. J. 217, 231–239. https://doi.org/10.1016/j.cej.2012.12.015 (2013).
    Google Scholar 
    Kazbar, A. et al. Effect of dissolved oxygen concentration on microalgal culture in photobioreactors. Algal Res. 39, 101432. https://doi.org/10.1016/j.algal.2019.101432 (2019).
    Google Scholar 
    Chisti, Y. Large-Scale Production of Algal Biomass: Raceway Ponds. in Algae Biotechnology: Products and Processes (eds Bux, F. & Chisti, Y.) 21–40 (2016). https://doi.org/10.1007/978-3-319-12334-9_2Sutherland, D. L., Howard-Williams, C., Turnbull, M. H., Broady, P. A. & Craggs, R. J. The effects of CO2 addition along a pH gradient on wastewater microalgal photo-physiology, biomass production and nutrient removal. Water Res. 70, 9–26. https://doi.org/10.1016/j.watres.2014.10.064 (2015).
    Google Scholar 
    Ketheesan, B. & Nirmalakhandan, N. Feasibility of microalgal cultivation in a pilot-scale airlift-driven raceway reactor. Bioresour. Technol. 108, 196–202. https://doi.org/10.1016/j.biortech.2011.12.146 (2012).
    Google Scholar 
    Sun, Y. et al. Boosting Nannochloropsis oculata growth and lipid accumulation in a lab-scale open raceway pond characterized by improved light distributions employing built-in planar waveguide modules. Bioresour. Technol. 249, 880–889. https://doi.org/10.1016/j.biortech.2017.11.013 (2018).
    Google Scholar 
    Posadas, E., Morales, M. M., Gomez, C., Acién, F. G. & Muñoz, R. Influence of pH and CO2 source on the performance of microalgae-based secondary domestic wastewater treatment in outdoors pilot raceways. Chem. Eng. J. 265, 239–248. https://doi.org/10.1016/j.cej.2014.12.059 (2015).
    Google Scholar 
    Ledda, C., Romero Villegas, G. I., Adani, F. & Acién Fernández, F. G. Molina Grima, E. Utilization of centrate from wastewater treatment for the outdoor production of Nannochloropsis Gaditana biomass at pilot-scale. Algal Res. 12, 17–25. https://doi.org/10.1016/j.algal.2015.08.002 (2015).
    Google Scholar 
    Huang, J. et al. Investigation on the performance of raceway ponds with internal structures by the means of CFD simulations and experiments. Algal Res. 10, 64–71. https://doi.org/10.1016/j.algal.2015.04.012 (2015).
    Google Scholar 
    Friedrich, K. et al. Reservoir Evaporation in the Western United States: Current Science, Challenges, and Future Needs 99, (2018). https://doi.org/10.1175/BAMS-D-15-00224.1Carrier, O. et al. Evaporation of water: evaporation rate and collective effects. J. Fluid Mech. 798, 774–786. https://doi.org/10.1017/jfm.2016.356 (2016).
    Google Scholar 
    Cuello, M. C., Cosgrove, J. J., Randhir, A., Vadiveloo, A. & Moheimani, N. R. Comparison of continuous and day time only mixing on tetraselmis Suecica (Chlorophyta) in outdoor raceway ponds. J. Appl. Phycol. 27, 1783–1791. https://doi.org/10.1007/s10811-014-0420-5 (2015).
    Google Scholar 
    Trainor, F. R. Reproduction in scenedesmus. Algae (The Korean J. Phycology). 11 (2), 183–201 (1996).
    Google Scholar 
    Barsanti, L. & Gualtieri, P. Algae: Anatomy, Biochemistry, and Biotechnology, Second Edition. (2014). https://doi.org/10.1201/b16544Rayen, F., Behnam, T. & Dominique, P. Optimization of a raceway pond system for wastewater treatment: a review. Crit. Rev. Biotechnol. 39, 422–435. https://doi.org/10.1080/07388551.2019.1571007 (2019).
    Google Scholar 
    Chuka-ogwude, D. et al. Effect of medium recycling, culture depth, and mixing duration on D. salina growth. Algal Res. 60, 102495. https://doi.org/10.1016/j.algal.2021.102495 (2021).
    Google Scholar 
    Zhu, L. et al. Nutrient removal and biodiesel production by integration of freshwater algae cultivation with piggery wastewater treatment. Water Res. 47, 4294–4302. https://doi.org/10.1016/j.watres.2013.05.004 (2013).
    Google Scholar 
    Marra, J., Bidigare, R. R. & Dickey, T. D. Nutrients and mixing, chlorophyll and phytoplankton growth. Deep Sea Res. Part. Oceanogr. Res. Papers. 37, 127–143. https://doi.org/10.1016/0198-0149(90)90032-Q (1990).
    Google Scholar 
    Lachmann, S. C., Mettler-Altmann, T., Wacker, A. & Spijkerman, E. Nitrate or ammonium: influences of nitrogen source on the physiology of a green Alga. Ecol. Evol. 9, 1070–1082. https://doi.org/10.1002/ece3.4790 (2019).
    Google Scholar 
    Jiang, R. et al. The joint effect of ammonium and pH on the growth of Chlorella vulgaris and ammonium removal in artificial liquid digestate. Bioresour. Technol. 325, 124690. https://doi.org/10.1016/j.biortech.2021.124690 (2021).
    Google Scholar 
    Tan, F. et al. Nitrogen and phosphorus removal coupled with carbohydrate production by five microalgae cultures cultivated in biogas slurry. Bioresour. Technol. 221, 385–393. https://doi.org/10.1016/j.biortech.2016.09.030 (2016).
    Google Scholar 
    Beuckels, A., Smolders, E. & Muylaert, K. Nitrogen availability influences phosphorus removal in microalgae-based wastewater treatment. Water Res. 77, 98–106. https://doi.org/10.1016/j.watres.2015.03.018 (2015).
    Google Scholar 
    Pena, A. C. C., Agustini, C. B., Trierweiler, L. F. & Gutterres, M. Influence of period light on cultivation of microalgae consortium for the treatment of tannery wastewaters from leather finishing stage. J. Clean. Prod. 263, 121618. https://doi.org/10.1016/j.jclepro.2020.121618 (2020).
    Google Scholar 
    Barboza-Rodríguez, R., Rodríguez-Jasso, R. M., Rosero-Chasoy, G., Rosales Aguado, M. L. & Ruiz, H. A. Photobioreactor configurations in cultivating microalgae biomass for biorefinery. Bioresour. Technol. 394, 130208. https://doi.org/10.1016/j.biortech.2023.130208 (2024).
    Google Scholar 
    Li, Y., Zhang, Q., Wang, Z., Wu, X. & Cong, W. Evaluation of power consumption of paddle wheel in an open raceway pond. Bioprocess. Biosyst Eng. 37, 1325–1336. https://doi.org/10.1007/s00449-013-1103-3 (2014).
    Google Scholar 
    Yadala, S. & Cremaschi, S. A. Dynamic Optimization Model for Designing Open-Channel Raceway Ponds for Batch Production of Algal Biomass. Processes 4, 10 (2016). https://doi.org/10.3390/pr4020010El-Chaghaby, A., Rashad, G., Abdel-Kader, S. F. & Rawash, S. A. Abdul Moneem, M. Assessment of phytochemical components, proximate composition and antioxidant properties of scenedesmus obliquus, chlorella vulgaris and spirulina platensis algae extracts. Egypt. J. Aquat. Biology Fisheries. 23, 521–526. https://doi.org/10.21608/ejabf.2019.57884 (2019).
    Google Scholar 
    Chen, M., Chen, Y. & Zhang, Q. A. Review of energy consumption in the acquisition of Bio-Feedstock for microalgae biofuel production. Sustainability 13, 8873. https://doi.org/10.3390/su13168873 (2021).
    Google Scholar 
    Bhatt, P. et al. Algae in wastewater treatment, mechanism, and application of biomass for production of value-added product. Environ. Pollut. 309, 119688. https://doi.org/10.1016/j.envpol.2022.119688 (2022).
    Google Scholar 
    Amorim, M. L., Soares, J., Vieira, B. B., Batista-Silva, W. & Martins, M. A. Extraction of proteins from the microalga scenedesmus obliquus BR003 followed by lipid extraction of the wet deproteinized biomass using hexane and Ethyl acetate. Bioresour Technol. 307, 123190. https://doi.org/10.1016/j.biortech.2020.123190 (2020).
    Google Scholar 
    Patnaik, R., Singh, N. K., Bagchi, S. K., Rao, P. S. & Mallick, N. Utilization of scenedesmus obliquus protein as a replacement of the commercially available fish meal under an algal refinery approach. Front. Microbiol. 10 https://doi.org/10.3389/fmicb.2019.02114 (2019).Olsen, M. F. L. et al. Outdoor cultivation of a novel isolate of the microalgae scenedesmus sp. and the evaluation of its potential as a novel protein crop. Physiol. Plant. 173 (2), 483–494. https://doi.org/10.1111/ppl.13532 (2021).
    Google Scholar 
    Skifa, I., Chauchat, N., Cocquet, P. H. & Guer, Y. L. Microalgae cultivation in raceway ponds: Advances, challenges, and hydrodynamic considerations. EFB Bioeconomy J. 5, 100073. https://doi.org/10.1016/j.bioeco.2024.100073 (2025).
    Google Scholar 
    Fakher, S., Khlaifat, A., Hossain, M. E. & Nameer, H. Rigorous review of electrical submersible pump failure mechanisms and their mitigation measures. J. Petroleum Explor. Prod. Technol. 11, 1507–1525. https://doi.org/10.1007/s13202-021-01271-6 (2021).
    Google Scholar 
    Zhang, Q. et al. Installation of flow deflectors and wing baffles to reduce dead zone and enhance flashing light effect in an open raceway pond. Bioresour Technol. 198, 150–156. https://doi.org/10.1016/j.biortech.2015.08.144 (2015).
    Google Scholar 
    Hoeniges, J., Zhu, K., Pruvost, J. & Legrand, J. Impact of Dropwise condensation on the biomass production rate in covered raceway ponds. Energies 14, 268 (2021).
    Google Scholar 
    Download referencesFundingThis research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2025R589), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.Author informationAuthors and AffiliationsInstitute of Environmental Sciences and Engineering (IESE), School of Civil and Environmental Engineering (SCEE), National University of Sciences and Technology (NUST), Islamabad, 44000, PakistanMuhammad Umer Abbas, Rashid Iftikhar, Nadeem Ullah & Faras Ahmad ShahbazDepartment of Sciences, School of Interdisciplinary Engineering & Science (SINES), National University of Science and Technology (NUST), Islamabad, PakistanSahar SaleemDepartment of Teaching and Learning, College of Education and Human Development, Princess Nourah Bint Abdulrahman University, Riyadh, 11671, Saudi ArabiaSarah Bader Alotaibi & Mashael M. AlfgehInstitute of Water Resources and Water Supply, Hamburg University of Technology (TUHH), Am Schwarzenberg-Campus 3, 21073, Hamburg, GermanyMuhammad Ali InamDepartment of Chemistry “Giacomo Ciamician”, University of Bologna, Via Selmi 2, Bologna, 40126, ItalyAhmad AakashAuthorsMuhammad Umer AbbasView author publicationsSearch author on:PubMed Google ScholarRashid IftikharView author publicationsSearch author on:PubMed Google ScholarSahar SaleemView author publicationsSearch author on:PubMed Google ScholarSarah Bader AlotaibiView author publicationsSearch author on:PubMed Google ScholarNadeem UllahView author publicationsSearch author on:PubMed Google ScholarMashael M. AlfgehView author publicationsSearch author on:PubMed Google ScholarMuhammad Ali InamView author publicationsSearch author on:PubMed Google ScholarFaras Ahmad ShahbazView author publicationsSearch author on:PubMed Google ScholarAhmad AakashView author publicationsSearch author on:PubMed Google ScholarContributionsMuhammad Umer Abbas: Writing—original draft, Writing—review & editing, Methodology, Validation, Conceptualization. Rashid Iftikhar: Writing—review & editing, Conceptualization, Validation, Project administration, Funding acquisition, Supervision. Sahar Saleem: Writing—review & editing and Methodology. Sarah Bader Alotaibi: Funding, resources & review. Nadeem Ullah: Resources and Review. Mashael M. Alfge: Funding & review. Muhammad Ali Inam: Methodology, Supervision, Validation, Investigation, final approval. Faras Ahmad Shahbaz: Writing—review & editing. Ahmad Aakash: Writing—review & editing.Corresponding authorCorrespondence to
    Rashid Iftikhar.Ethics declarations

    Competing interests
    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    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 articleAbbas, M.U., Iftikhar, R., Saleem, S. et al. Impact of mixing duration on growth and nutrient removal efficiency of Scenedesmus sp. in a novel raceway pond system.
    Sci Rep (2025). https://doi.org/10.1038/s41598-025-31982-3Download citationReceived: 08 August 2025Accepted: 06 December 2025Published: 15 December 2025DOI: https://doi.org/10.1038/s41598-025-31982-3Share 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
    KeywordsNutrient recovery; evaporation loss; energy efficiencySubmersible pumpReactor designHydrodynamics More

  • in

    Functional trait variations of the invasive plant Alternanthera Philoxeroides and the native plant Ludwigia peploides under nitrogen addition

    AbstractQuantifying the variation and coordination patterns of plant functional traits across different organs under environmental changes is crucial for understanding plant invasion and adaptation mechanisms. This study employed a space-for-time substitution experiment to compare the differential responses of root and leaf functional traits and their coordination in the invasive plant Alternanthera philoxeroides and the native plant Ludwigia peploides to nitrogen addition during different invasion degree. The results showed that: (1) nitrogen addition promoted the growth of both species, with A. philoxeroides exhibited greater biomass sensitivity. Compared to the positive effects of nitrogen fertilization, nitrogen addition facilitated A. philoxeroides in displacing L. peploides in communities with 50% (2 A. philoxeroides seedlings) and 75% (3 A. philoxeroides seedlings) invasion degree. (2) invasion degree, nitrogen addition, and their interaction significantly influenced most root and leaf traits of both species. But the two species differed markedly in their response of root and leaf traits to environmental factors. (3) The correlations between root traits, leaf traits, and total biomass were stronger in A. philoxeroides than in L. peploides, as were the linkages between root and leaf traits. Under environmental changes, the two species exhibited distinct adaptive strategies in root and leaf traits, with A. philoxeroides’s trait advantages likely contributing to its invasion success. In conclusion, our study demonstrates that nitrogen deposition facilitates alien plant invasion, particularly in mixed communities experiencing moderate to severe invasion.

    Similar content being viewed by others

    Identify potential allelochemicals from Humulus scandens (Lour.) Merr. root extracts that induce allelopathy on Alternanthera philoxeroides (Mart.) Griseb.

    Article
    Open access
    29 March 2021

    Effects of arbuscular mycorrhizal fungi and soil substrate on invasive plant Alternanthera philoxeroides

    Article
    Open access
    01 July 2025

    Leaf economic strategies drive global variation in phosphorus stimulation of terrestrial plant production

    Article
    Open access
    01 July 2025

    IntroductionIn natural environments, native plant communities often face varying degrees of invasion pressure from exotic plants1,2,3,4. During the establishment phase, invasive plants must overcome resistance from native vegetation, yet they often succeed in occupying niches due to their superior competitive abilities4. In the expansion phase, competitive advantages directly determine the degree of dominance by invasive species and the severity of their ecological impacts, where intense interspecific competition may even lead to the exclusion of native species4. These competitive dynamics are fundamentally driven by differences in functional traits among species and are maintained through niche differentiation and mechanisms of competitive coexistence5,6,7.Plants optimize key physiological processes such as photosynthetic carbon acquisition, structural support, and nutrient uptake by adjusting biomass allocation patterns to organs like roots and leaves8. Variations in root and leaf morphology, along with tissue nutrient content (e.g., nitrogen concentration), directly reflect plant strategies and efficiency in resource utilization8,9,10. Furthermore, secondary metabolites serve as critical biochemical traits, not only providing defensive compounds but also significantly influencing plant adaptation to biotic and abiotic environments5,8. Non-structural carbohydrates (NSCs), as labile carbon reserves, play a central role in plant growth and environmental adaptation by supplying carbon skeletons for development, fueling respiratory metabolism, and participating in osmotic regulation, among other processes10. Previous research has shown that invasive plants often exhibit greater leaf mass fraction, specific root length (SRL), and root nitrogen concentration, as well as release higher amounts of allelochemicals, compared to co-occurring native species5,8,9,10.However, most existing studies in invasion ecology focus on either “pre-invasion” or “post-invasion” community states1,9,11,12,13,14, often overlooking the continuum of invasion degree and failing to systematically analyze how different invasion degrees differentially affect the functional traits of invasive and native plants. This knowledge gap hinders the identification of key invasive traits that determine competitive outcomes8,9,10.Moreover, as a critical component of global change, biological invasions often interact synergistically with anthropogenic drivers such as nitrogen deposition, climate warming, and rising atmospheric CO₂ levels15,16. Research indicates that the success of invasive plants in establishment and expansion closely depends on resource availability and plasticity in their functional traits12. In recent decades, global fossil fuel combustion and synthetic fertilizer use have led to a significant increase in atmospheric nitrogen deposition17,18. China has become the world’s third-largest nitrogen deposition hotspot after Europe and North America, with southern regions experiencing an annual deposition rate of up to 63.53 kg N ha⁻¹ yr⁻¹5,19. As a key nutrient for plant growth, nitrogen availability profoundly influences plant performance and adaptive strategies during invasion16,20,21,22,23,24,25. For instance, in high-nutrient microenvironments, invasive plants often enhance their competitiveness through rapid adjustments in root and leaf trait plasticity5. Nevertheless, there remains a scarcity of systematic research on the responses of above- and below-ground traits in invasive and native plants under varying nitrogen levels and invasion degree.Previous studies have found that aboveground and belowground components of plants are closely interconnected: plants compete for light through above-ground organs (e.g., leaves) while simultaneously vying for nutrients and water via below-ground structures (e.g., roots)4. Recent studies have highlighted the critical importance of coordinated root and leaf functional traits in responding to heterogeneous above- and below-ground resource availability26,27. Such coordination enhances a plant’s capacity to either optimize acquisition of limited resources or minimize demand for specific resources28. For instance, research has demonstrated systematic correlations between root morphological traits and leaf traits associated with nutrient utilization26,29. This reflects an integrated whole-plant strategy where above- and below-ground components function synergistically to adapt to environmental variables including temperature, light intensity, and nutrient/water availability30. Despite their well-established role in plant resource-use strategies31, current understanding of how inter-organ trait coordination responds to environmental changes remains fragmented.Our earlier study observed that, in a 60-day short-term experiment, although invasive and native plant communities exhibited differential responses in total and root biomass, root morphology, and exudate composition to nitrogen addition, neither showed a significant preference for ammonium vs. nitrate nitrogen8. Since leaves are central organs in above-ground resource competition4, the physiological and morphological responses of leaves to nitrogen addition were not sufficiently addressed in previous work. Therefore, this study systematically investigates the effects of nitrogen addition on whole-plant biomass, root and leaf morphology, and root exudates in the invasive species Alternanthera philoxeroides and the native species Ludwigia peploides.Based on the above background, this study establishes three nitrogen addition levels (i.e., control, low, and high nitrogen treatments) and five invasion scenarios (i.e., no invasion, early invasion, mid-invasion, dominant invasion, and native species migration period). We hypothesize that: (1) nitrogen addition will promote the growth of both A. philoxeroides and L. peploides, but increasing invasion degree will suppress the growth of L. peploides; (2) nitrogen addition and invasion degree will significantly alter key root and leaf functional traits in both species; (3) nitrogen addition and invasion degree can regulate the invasion success of A. philoxeroides by modulating root traits, leaf traits, and their interactions, whereas this pattern will not be observed in L. peploides.ResultVariation in total biomassBoth invasion degree and nitrogen level significantly affected the individual plant biomass of A. philoxeroides and L. peploides, although their interaction was not significant (Fig. 1). Under the same invasion degree, both nitrogen addition treatments significantly increased total individual biomass of A. philoxeroides compared to the control (Fig. 1a). But only high nitrogen significantly increased the total individual biomass of L. peploides (Fig. 1b). Under the same nitrogen treatment, the total biomass of A. philoxeroides showed an increasing trend with the degree of invasion, while the opposite was true for L. peploides (Fig. 1a, b).Fig. 1Effect of invasion degree and nitrogen addition on the total dry biomass of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. Three nitrogen addition levels: 0, 6, and 12 g N·m⁻²·yr⁻¹, designated as CK, low nitrogen (Low N), and high nitrogen (High N) treatments, respectively. Five invasion scenarios: no invasion (0%), early invasion (25%), mid-invasion (50%), dominant invasion (75%), and native species migration period (100%). Sample sizes: n = 60.Full size imageVariation in root traitsAs shown in Fig. 2, the invasion degree and nitrogen level, as well as their interaction, significantly affected all root trait indices of A. philoxeroides (except for RMF and SRL) and L. peploides (except for RD).For A. philoxeroides, under different invasion degree, both nitrogen application treatments tended to decrease RMF, RD, and SRL (at early invasion (25%) and mid-invasion (50%)) compared to the CK. In contrast, both nitrogen levels increased root starch, NSC, nitrogen content, and SRL (at dominant invasion (75%) and native species migration period (100%)) (Fig. 2a-f). For L. peploides, across invasion degree, RMF gradually decreased with increasing nitrogen application, while RD showed the opposite trend (Fig. 2g, h). High nitrogen significantly reduced SRL, whereas the effect of low nitrogen varied depending on invasion degree (Fig. 2i). Both nitrogen treatments significantly decreased root starch and NSC content but markedly increased root nitrogen content (Fig. 2j-l).Compared to the native species migration period (100%), the presence of L. peploides significantly increased root starch and NSC content in A. philoxeroides (Fig. 2d, e). Under the CK treatment, the presence of L. peploides significantly reduced RMF and RD but increased SRL in A. philoxeroides, while its effect on root nitrogen content was minimal (Fig. 2a-c, f). Under both nitrogen application levels, the presence of L. peploides significantly decreased RMF (at mid-invasion (50%) and dominant invasion (75%)) and increased SRL, root starch, and NSC content in A. philoxeroides (Fig. 2a, c-e). Under low nitrogen, the presence of L. peploides significantly reduced root nitrogen content but increased RD (Fig. 2b, f). Under high nitrogen, it significantly increased RD (at early invasion (25%) and mid-invasion (50%)) and root nitrogen content (at mid-invasion (50%) and dominant invasion (75%)) in A. philoxeroides (Fig. 2b, f).Compared to the absence of A. philoxeroides invasion, the presence of A. philoxeroides showed no significant effect on the RD of L. peploides (Fig. 2h). Under the CK conditions, A. philoxeroides presence exhibited a decreasing trend in RMF, root starch, NSC and nitrogen content of L. peploides, while no significant effect was observed on SRL (Fig. 2g, i-l). Under both nitrogen application treatments, A. philoxeroides presence had minimal effects on root starch and NSC content of L. peploides (Fig. 2j, k). With low nitrogen treatment, A. philoxeroides presence showed limited influence on RMF but significantly increased SRL and root nitrogen content of L. peploides (Fig. 2g, i, l). Under high nitrogen treatment, A. philoxeroides presence significantly reduced RMF of L. peploides while demonstrating minor effects on SRL and root nitrogen content (Fig. 2g, i, l).Changes in root-secreted secondary metabolitesThe results of the Adonis analysis revealed that invasion degree significantly affected the composition of root exudates in both plant species, whereas nitrogen application showed no significant effect on their root exudate profiles (Fig. 3a and c).For A. philoxeroides, compared to the native species migration period (100%), the presence of L. peploides significantly reduced phenolic compound content in root exudates and showed a decreasing trend for terpenoids (except under both nitrogen treatments) and alkanes content. In contrast, organic acids and amides exhibited opposite trends, while alkaloids content remained unaffected (Fig. 3b). For L. peploides, compared to the absence of A. philoxeroides invasion, the presence of A. philoxeroides showed minimal effects on the content of phenolic compounds, alkanes, alkaloids, organic acids (under both nitrogen treatments), and amides (under CK and low nitrogen treatments) in its root exudates. However, a decreasing trend was observed for terpenoids content, while an increasing trend was noted for organic acids (under CK treatment) and amides (under high nitrogen treatment) (Fig. 3d).The Pearson correlation analysis revealed significant relationships between root traits and total biomass for both species (Fig. 4). For A. philoxeroides, RD, SRL, and root-secreted amides content showed significant negative correlations with total biomass, whereas root nitrogen concentration, phenolics, and alkanes content in root exudates exhibited significant positive correlations (Fig. 4). For L. peploides, RMF, and root-secreted terpenoids content showed significant positive correlations with total biomass, whereas SRL, root starch and root-secreted organic acids content exhibited significant negative correlations (Fig. 4).Fig. 2Effect of invasion degree and nitrogen addition on the root traits of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. (a, g) root mass fraction; (b, h) root diameter; (c, i) specific root length (SRL); (d, j) root starch content; (e, k) root total non-structural carbohydrates content (Root NSC); (f, l) root nitrogen content. Three nitrogen addition levels: 0, 6, and 12 g N·m⁻²·yr⁻¹, designated as CK, low nitrogen (Low N), and high nitrogen (High N) treatments, respectively. Five invasion scenarios: no invasion (0%), early invasion (25%), mid-invasion (50%), dominant invasion (75%), and native species migration period (100%). Sample sizes: n = 60.Full size imageFig. 3Effect of invasion degree and nitrogen addition on the content of secondary metabolites produced by the roots of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. (a, c) show the principal component analysis (PCA) of secondary metabolites produced by the roots of A. philoxeroides and L. peploides over invasion degree and nitrogen addition gradient analyzed using gas chromatography-mass spectrometry (GC–MS); (b, d) displays the percentage distribution of secondary metabolites produced by the roots (alkaloids, alkanes, amides, organic acids, phenols and terpenes) of A. philoxeroides and L. peploides under different invasion degree and nitrogen level treatments. RPA(%) represents the percentage of secretory products relative to the total secondary metabolite content. Three nitrogen addition levels: 0, 6, and 12 g N·m⁻²·yr⁻¹, designated as CK (N0), low nitrogen (N6, Low N), and high nitrogen (N12, High N) treatments, respectively. Five invasion scenarios: no invasion (0%), early invasion (25%), mid-invasion (50%), dominant invasion (75%), and native species migration period (100%).Full size imageFig. 4Pearson correlation analysis between total biomass and root traits of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. RMF represents root mass fraction; RD represents root diameter; SRL represents specific root length; Root NSC represents total root non-structural carbohydrates content; and Root N represents root nitrogen concentration.Full size imageChanges in leaf traitsAs shown in Fig. 5, invasion degree, nitrogen level, and their interaction significantly affected all leaf trait indices of A. philoxeroides (except LMF and SLA) and L. peploides (except SLA).Compared to the CK, nitrogen addition exhibited an increasing trend in LMF for both A. philoxeroides and L. peploides, with a more pronounced effect in A. philoxeroides (Fig. 5a, f). Both nitrogen levels significantly enhanced leaf nitrogen concentration in both species (Fig. 5e, j) but markedly reduced leaf starch (Fig. 5c, h) and NSC content (Fig. 5d, i). Differently, nitrogen application showed a trend of reducing the SLA of A. philoxeroides, but significantly increased the SLA of L. peploides (Fig. 5b, g).Compared with the native species migration period (100%), the presence of L. peploides non-significant effects the SLA of A. philoxeroides, and had a moderate effect on the LMF under nitrogen treatment. However, it significantly increased the leaf starch and NSC content under CK treatment, and significantly reduced the LMF under CK treatment (at early invasion (25%) and mid-invasion (50%)) and NSC content and leaf nitrogen content under high nitrogen treatment (Fig. 5a-e). Compared to the absence of A. philoxeroides invasion, the presence of A. philoxeroides had minimal effects on LMF and leaf nitrogen content of L. peploides, but significantly reduces its SLA (except at early invasion (25%)), the starch and NSC content in its leaves vary depending on the invasion degree of A. philoxeroides (Fig. 5f-j).Changes in leaf-secreted secondary metabolitesResults from the Adonis analysis revealed that invasion degree, nitrogen level, and their interaction significantly influenced the leaf exudate composition of A. philoxeroides (Fig. 6a). In contrast, only nitrogen level exhibited a significant effect on the leaf exudate profile of L. peploides (Fig. 6c).For A. philoxeroides, nitrogen application showed an increasing trend in leaf terpenoids content compared to the CK, while demonstrating decreasing trends for organic acids and alkaloids (under high nitrogen treatment). No significant effects were observed on phenols, amides, alkanes, or alkaloids (under low nitrogen treatment) (Fig. 6b). In contrast, for L. peploides, nitrogen application had no significant effect on leaf terpenoids, amides, and alkanes relative to CK. Minor effects were observed on phenols and alkaloids, while an increasing trend was noted for organic acids content (Fig. 6d).For A. philoxeroides, compared to the native species migration period (100%), the presence of L. peploides exhibited minimal effects on leaf terpenoids and amides content, no significant effects on leaf phenols and alkaloids content (under both CK and low nitrogen treatments), a decreasing trend in organic acids content, but significant increases in alkanes content, significant reductions in alkaloids content (under high nitrogen treatment) (Fig. 6b).The Pearson correlation analysis revealed that the biomass of A. philoxeroides was significantly positively correlated with its LMF, leaf nitrogen concentration, and terpenoids content secreted by leaves, whereas its SLA, leaf starch, NSC, and alkanes, amides secreted by leaves showed a significant negative correlation with total biomass (Fig. 7). For L. peploides, its individual biomass was significantly positively correlated with its SLA, leaf nitrogen concentration, and terpenoids content secreted by its leaves, whereas its leaf starch, NSC, and phenols, amides secreted by leaves showed a significant negative correlation with total biomass (Fig. 7).Fig. 5Effect of invasion degree and nitrogen addition on the leaf traits of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. (a, f) leaf mass fraction; (b, g) specific leaf area (SLA); (c, h) leaf starch content; (d, i) leaf total non-structural carbohydrates content (Leaf NSC); (e, j) leaf nitrogen content. Three nitrogen addition levels: 0, 6, and 12 g N·m⁻²·yr⁻¹, designated as CK, low nitrogen (Low N), and high nitrogen (High N) treatments, respectively. Five invasion scenarios: no invasion (0%), early invasion (25%), mid-invasion (50%), dominant invasion (75%), and native species migration period (100%). Sample sizes: n = 60.Full size imageFig. 6Effect of invasion degree and nitrogen addition on the content of secondary metabolites produced by the leaves of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. (a, c) show the principal component analysis (PCA) of secondary metabolites produced by the leaves of A. philoxeroides and L. peploides over invasion degree and nitrogen addition gradient analyzed using gas chromatography–mass spectrometry (GC–MS); (b, d) displays the percentage distribution of secondary metabolites produced by the leaves (alkaloids, alkanes, amides, organic acids, phenols and terpenes) of A. philoxeroides and L. peploides under different invasion degree and nitrogen level treatments. RPA(%) represents the percentage of secretory products relative to the total secondary metabolite content. Three nitrogen addition levels: 0, 6, and 12 g N·m⁻²·yr⁻¹, designated as CK (N0), low nitrogen (N6, Low N), and high nitrogen (N12, High N) treatments, respectively. Five invasion scenarios: no invasion (0%), early invasion (25%), mid-invasion (50%), dominant invasion (75%), and native species migration period (100%).Full size imageFig. 7Pearson correlation analysis between total biomass and leaf traits of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. LMF represents leaf mass fraction; SLA represents specific leaf area; Leaf NSC represents total leaf non-structural carbohydrates content; and leaf N represents leaf nitrogen concentration.Full size imageCorrelations among total biomass, root traits, and leaf traitsA. philoxeroides exhibited 56 significant root-leaf trait correlations, whereas L. peploides showed 50 significant root-leaf trait correlations (Fig. 8a, b). The PLS-PM model showed that nitrogen application level (path coefficient = 0.340, P < 0.001) and invasion degree (path coefficient = 0.249, P < 0.05) had a significant positive effect on the total biomass of A. philoxeroides, while nitrogen application level (P < 0.001) and invasion degree (P < 0.001) had a significant negative effect on the root traits of A. philoxeroides; At the same time, nitrogen application level (P < 0.001) and invasion degree (P < 0.05) also had a significant negative effect on the leaf traits of A. philoxeroides, but subsequently, changes in leaf traits promoted the growth of A. philoxeroides (Fig. 9a). There was a significant negative correlation between the invasion degree (P < 0.001) and the total biomass of L. peploides; There was also a significant negative correlation between nitrogen application level and its root traits (P < 0.001) and leaf traits (P < 0.001), but there was a significant positive correlation between invasion degree and its root traits (P < 0.001) and leaf traits (P < 0.01) (Fig. 9b).Fig. 8Pearson correlation analysis between root traits and leaf traits of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides under different invasion degree and nitrogen addition. RMF and LMF: represent root/leaf mass fraction; RD represent root diameter; SRL represent specific root length; SLA represent specific leaf area; RS and LS: represent root/leaf starch content; RNSC and LNSC: represent root/leaf total non-structural carbohydrate content; RNC and LNC: represent root/leaf nitrogen content.Full size imageFig. 9The partial least squares path model of invasive plant Alternanthera philoxeroides and native plant Ludwigia peploides. The Partial Least Squares Path Model illustrates the influence of nitrogen level and invasion degree on total biomass of A. philoxeroides (a) and L. peploides (b) by modulating their root traits and leaf traits. Arrow widths represent the strength of the path coefficient, and solid lines indicate significant correlations. Red and blue lines denote positive and negative pathways, respectively. Significance levels are denoted by asterisks: *p < 0.05, **p < 0.01, ***p < 0.001. Reflective latent variables (orange blocks) are indicated by the measured variables (blue blocks), with their respective weights shown. RMF and LMF: represent root/leaf mass fraction; RD represent root diameter; SRL represent specific root length; SLA represent specific leaf area; Root NSC and Leaf NSC: represent root/leaf total non-structural carbohydrate content; Root N and Leaf N: represent root/leaf nitrogen content.Full size imageDiscussionWhile the 90-day experimental period cannot fully capture the long-term dynamics of invasion processes, the short-term response patterns observed in this study remain highly significant. These early adaptive changes may represent critical first steps toward successful invasion.Nitrogen addition was likely to facilitate the displacement of L. peploides by A. philoxeroides in plant communities experiencing mid-invasion (50%) and dominant invasion (75%)Biomass serves as a direct measure of plant growth performance and resource-use efficiency28,30. Nitrogen enrichment generally promotes biomass production of invasive species and native species, invasive plants typically exhibit greater sensitivity to nitrogen enrichment compared to their native counterparts24,28,32,33,34, a pattern consistently observed in our study (Fig. 1a, b). As a critical nutrient regulating plant growth and productivity, nitrogen availability often determines the degree of biomass accumulation by fulfilling metabolic demands16,21,25,35. Compared to the native L. peploides, A. philoxeroides was a fast-growing, opportunistic perennial herb that typically exhibits faster phenotypic adjustments to nutrient addition or shifts in nutrient availability13.Notably, under nitrogen addition, A. philoxeroides outproduced L. peploides in biomass accumulation within mid-invasion (50%) and dominant invasion (75%) communities (Fig. 1a, b). In coexisting communities, A. philoxeroides and L. peploides occupy analogous ecological niches, and the high overlap of ecological niches may lead to fierce competition between invasive species and closely related native species occupying the same domain range5. Previous studies have shown that invasive plants can suppress native species by forming high-density canopies that intercept light, fiercely compete for nutrients, and even secrete allelopathic substances2. Therefore, as the relative abundance of A. philoxeroides increases, its ability to compete for space and nutrients in the community becomes stronger, and the growth of most native plants is inhibited, resulting in a decrease in individual biomass12. Atmospheric nitrogen deposition levels have increased significantly in recent years (cite source), with demonstrated impacts on plant communities in highly invaded ecosystems36. Our findings reveal that the early-stage invasive L. peploides exhibited superior growth performance compared to A. philoxeroides under these conditions (Fig. 1a, b). This temporal advantage suggests that early detection and intervention may represent the most effective strategy for mitigating A. philoxeroides invasion, particularly in the context of rising nitrogen availability.Variation in root traitsPlant roots, as critical organs for soil nutrient absorption, play a vital role in mediating plant-plant interactions and acquiring soil resources27. Their morphological and architectural traits can also predict a plant’s ability to tolerate competitors5. In this experiment, A. philoxeroides exhibited a higher SRL (Fig. 2c, i). Compared to the CK, nitrogen application tended to reduce its RD, whereas low nitrogen had no effect on L. peploides RD, and high nitrogen even increased it (Fig. 2b, h). Generally, thinner roots (but with higher SRL) may possess narrower water-conducting vessels and faster turnover rates, enhancing soil exploration and exploitation efficiency37,38. This trait reduces the cost of root proliferation in fluctuating environments, thereby improving nutrient absorption efficiency37,38.Differences in root trait changes were also observed between the two plant species at different invasion stages. In this experiment, compared to the native species migration period (100%), the presence of L. peploides had a minor effect on the root nitrogen content of A. philoxeroides under the CK treatment. However, under low nitrogen treatment, L. peploides significantly reduced the root nitrogen content of A. philoxeroides, while a slight increasing trend was observed under high nitrogen treatment (Fig. 2f). Conversely, compared to plots without A. philoxeroides invasion, the presence of A. philoxeroides tended to decrease the root nitrogen content of L. peploides under CK treatment, significantly increased it under low nitrogen treatment, and had a negligible effect under high nitrogen treatment (Fig. 2l). These patterns may be linked to changes in root exudate composition between the two species. Generally, higher root nutrient concentrations may correlate with greater abundance of soil-borne pathogens and root-feeding insects39,40. Consequently, plants may allocate additional resources to defense structures, including the secretion of secondary metabolites into roots for chemical protection41. Among the secondary metabolites detected in root exudates, we observed that the presence of L. peploides tended to elevate the secretion of organic acids and amides by A. philoxeroides roots compared to the native species migration period (100%) (Fig. 3b). In contrast, the presence of A. philoxeroides had minimal effects on the composition of L. peploides root exudates (Fig. 3d). Given that organic acids and amides are known allelochemicals which can affect the growth of neighboring plants8. Our finding suggests a potential feedback mechanism: the presence of L. peploides induces A. philoxeroides to release such compounds, likely to negatively impacting the growth of the native species itself.Further, physiological traits serve as sensitive indicators of plant environmental responses, typically detectable earlier than morphological traits42. Among these, non-structural carbohydrates (NSC, including soluble sugars and starch) availability profoundly influences plant growth and long-term survival43,44,45,46. In this experiment, nitrogen application and the presence of L. peploides significantly increased starch and NSC content in A. philoxeroides roots compared to both the CK and native species migration period (100%) (Fig. 2d, e). In contrast, nitrogen fertilization significantly reduced starch and NSC content in L. peploides roots relative to CK. Furthermore, compared to plots without A. philoxeroides invasion, the invasive species’ presence showed a tendency to decrease starch and NSC content in L. peploides roots under CK treatment, with minimal effects under nitrogen addition (Fig. 2j, k). These findings suggest divergent resource allocation strategies: A. philoxeroides allocates substantial resources to root storage, while L. peploides prioritizes rapid growth.Variation in leaf traitsLeaves, as the primary photosynthetic organs of plants, exhibit higher metabolic activity than roots and stems44. Typically, increased leaf biomass exerts strong control over aboveground resource acquisition47. Traits such as high SLA and elevated leaf nitrogen content were generally associated with enhanced photosynthetic capacity25,37,48. Additionally, the mobilization of starch and NSC stored in leaves can increase leaf respiration rates, thereby meeting the elevated carbohydrate demands of maintenance respiration44. Collectively, these adaptive traits improve a plant’s ability to absorb and utilize resources under changing environmental conditions, ultimately supporting greater aboveground growth. In this study, we found that nitrogen application enhanced the photosynthetic capacity of both study species compared to the CK, as evidenced by increased LMF, higher leaf nitrogen content, and reduced starch and NSC concentrations in leaves (Fig. 5). These results align with prior research demonstrating that nitrogen fertilization generally promotes leaf photosynthesis and plant growth44.We observed distinct patterns in leaf trait modifications of A. philoxeroides compared to L. peploides under varying invasion scenarios. Generally, elevated leaf nitrogen content alters plant interactions with nutrient-rich organisms and increases palatability to herbivores39,40. In our experiments, the presence of L. peploides significantly reduced leaf nitrogen content in A. philoxeroides relative to the native species migration period (100%) (Fig. 5e). Conversely, A. philoxeroides invasion showed minimal effects on L. peploides leaf nitrogen content (Fig. 5j). These differential responses may reflect variations in leaf exudate composition between the species. Compared to the native species migration period (100%), L. peploides presence exerted limited effects on terpenoids, amides, phenols, and alkaloids (under CK and low nitrogen treatments) in A. philoxeroides leaves, while showing a tendency to reduce organic acids and significantly decreasing alkaloids content under high nitrogen treatment (Fig. 6b). Notably, A. philoxeroides invasion did not alter the exudate profile of L. peploides leaves (Fig. 6d). The observed reduction in leaf nitrogen content, coupled with limited allelochemical secretion, we hypothesize that an adaptive strategy in A. philoxeroides to minimize herbivory pressure.The correlation between root and leaf functional traits and their adaptive strategiesTrait correlations were considered to reflect either trade-offs or synergistic optimization in resource allocation to meet fundamental survival requirements38,49. Plants exhibiting stronger root-leaf trait coordination may demonstrate greater growth success and survival when exposed to environmental variability30,50. This study revealed that under different nitrogen addition and invasion degree treatments, most root-leaf traits of both plant species exhibited significant correlations, though the strength and patterns of these correlations differed markedly. Compared to the native species L. peploides (involving 50 root-leaf trait correlations), A. philoxeroides demonstrated stronger root-leaf trait integration (involving 56 root-leaf trait correlations) (Fig. 8). This divergent trait correlation pattern may confer important adaptive value: it not only helps plants minimize negative impacts in unfavorable environments but also effectively enhances their capacity for survival, growth, and reproduction30. The differences in trait correlation strategies between A. philoxeroides and L. peploides reflect the diverse manifestations of ecological trait plasticity during environmental adaptation among different species.Results of the Partial Least Squares Path Modeling (PLS-PM) revealed that nitrogen addition and invasion degree can modulate the growth of A. philoxeroides through regulating root traits, leaf traits, and their interactions – a pattern not observed in L. peploides (Fig. 9a, b). Specifically, nitrogen addition and invasion degree not only directly increased the per-plant biomass of A. philoxeroides, but also indirectly promoted its biomass accumulation by altering leaf traits—such as increasing the leaf mass fraction and leaf nitrogen content, while reducing leaf starch content (Fig. 9a). In contrast, nitrogen addition had no significant effect on the per-plant biomass of L. peploides, whereas its biomass significantly decreased with increasing invasion degree. Furthermore, under conditions of nitrogen addition and invasion degree, the root traits, leaf traits, and their interrelationships in L. peploides did not exhibit significant regulatory effects on its total biomass (Fig. 9b). Previous research has indicated that trait plasticity and trait coordination play crucial roles in plant adaptation to environmental changes and may facilitate niche expansion30. Our findings further corroborate that, compared to the native species L. peploides, A. philoxeroides exhibits greater phenotypic plasticity in response to environmental variation. This advantage was likely attributable to its more tightly coordinated root-leaf trait relationships. This enhanced trait integration may be a key mechanism underlying the ecological success of this invasive species in heterogeneous habitats.In summary, a critical invasion mechanism of successful alien species lies in their superior trait values compared to co-occurring native species, enabling them to outcompete native flora and facilitate establishment in recipient habitats8,28. Notably, variations in root and leaf traits result from complex interactions between multiple biotic and abiotic factors26,51. Consequently, comprehensive quantification of interspecific differences in these traits is essential for elucidating the mechanisms underlying plant invasions.ConclusionIn conclusion, we observed that A. philoxeroides and L. peploides displayed differing belowground and aboveground trait responses under varying invasion scenarios and nitrogen treatments. Nitrogen addition promoted growth in both species, whereas invasion pressure had a more pronounced negative effect on L. peploides. These species-specific patterns appear linked to differential adjustments in root and leaf trait expression. Overall, compared to the native plant L. peploides, the invasive plant A. philoxeroides had more advantages in root and leaf traits under environmental treatment, and the correlation between root and leaf traits was stronger. These findings suggest that under progressive atmospheric nitrogen deposition, A. philoxeroides may progressively displace L. peploides, particularly in communities experiencing mid-invasion (50%) to dominant invasion (75%) invasion degrees. However, the limited soil types and plant species used in this study constrain the generalizability of our findings. Future experimental designs should incorporate more naturalistic scenarios to enhance ecological relevance.Materials and methodsStudy speciesThis study selected the invasive plant Alternanthera philoxeroides and the native plant Ludwigia peploides as research subjects. A. philoxeroides and L. peploides both exhibit rapid expansion capability through clonal growth and can quickly adapt to environmental changes by adjusting their above- and below-ground traits46,52. In natural ecosystems, these two species often co-occur over broad geographical ranges, sharing similar habitat types such as rice paddies, wetlands, canals, ponds, and ditches52. A. philoxeroides, native to South America, was now widely distributed across many regions worldwide and has become one of the most aggressive invasive alien species in China, causing significant ecological and economic impacts in China and numerous other countries46. In contrast, L. peploides is native to Zhejiang, Fujian, and eastern Guangdong in China and serves as a dominant native species in subtropical to tropical regions of the country52.Experimental designIn April 2022, 300 seedlings of A. philoxeroides and 300 seedlings of L. peploides were collected at the Liangzihu national field ecological research station of Wuhan University (N30°05–30°18, E114°21–114°39). They were then cultured in a aquarium tanks (100 × 30 × 50 cm, L × W × H) filled with 30 cm deep lake sediment (TC, 31.22 mg·g⁻¹; TN, 4.09 mg·g⁻¹; TP, 2.27 mg·g⁻¹). These two types of clonal plants were grown under greenhouse conditions (The mean annual temperature is 25 °C with an average annual sunshine duration of 1,810 h) for one year.On May 1, 2023, we selected 150 ramets each of A. philoxeroides and L. peploides from the pre-cultured seedlings, choosing individuals with biomass (approximately 1.9 g) and height (approximately 15 cm). These selected plants were then transplanted into stainless steel pots (70 cm inner diameter × 20 cm height) according to experimental treatments. Each pot was filled with approximately 15 cm of lake sediment (TC: 31.22 mg·g⁻¹; TN: 4.09 mg·g⁻¹; TP: 2.27 mg·g⁻¹). The experimental pots were randomly arranged on the outdoor cement platform at Liangzi Lake Ecological Station, which featured an open, flat terrain without obstructions and received ample sunlight.Following established methodologies14, we employed a space-for-time substitution approach to simulate the progressive invasion process of A. philoxeroides. Five invasion scenarios were established: the total biomass of 4 plants per pot was strictly controlled within the range of 7.40–7.60 g. (1) no invasion period (0%): composed exclusively of 4 L. peploides seedlings; (2) early invasion period (25%): consisting of 1 A. philoxeroides and 3 L. peploides seedlings; (3) mid-invasion period (50%): containing 2 A. philoxeroides and 2 L. peploides seedlings; (4) dominant invasion period (75%): comprising 3 A. philoxeroides and 1 L. peploides seedling; (5) native species migration period (100%): represented by 4 A. philoxeroides seedlings. Throughout the experiment, any spontaneously occurring rare weeds in the pots were manually removed. Furthermore, according to previous research1,8,16,23, each invasion treatment was coupled with three nitrogen addition levels: 0, 6, and 12 g N·m⁻²·yr⁻¹, designated as N0 (control, CK), N6 (low nitrogen), and N12 (high nitrogen) treatments, respectively. The N6 level represents the current average nitrogen deposition rate recorded in certain regions of China, while N12 corresponds to potential future high nitrogen deposition scenarios1,16,23. Starting in May 7, 2023, artificial nitrogen addition (NH4NO3) to the pots on a weekly basis according to the designated nitrogen deposition levels. The experiment was conducted for 90 days, concluding on August 5, 2023. The complete experimental design was illustrated in Fig. 10. Although the space-for-time substitution approach has inherent limitations, this study implemented strict controls to ensure consistent total biomass and plant numbers per pot, which helps partially compensate for the constraints of short-term observations.Fig. 10Illustrates the experimental design and provides a visual representation of the experiment. Five invasion scenarios were established: native species migration period (100%), 4 A. philoxeroides seedlings; dominant invasion period (75%), 3 A. philoxeroides and 1 L. peploides seedling; mid-invasion period (50%), 2 A. philoxeroides and 2 L. peploides seedlings; early invasion period (25%), 1 A. philoxeroides and 3 L. peploides seedlings; no invasion period (0%), 4 L. peploides seedlings.Full size imageSampling collect and trait measurementIn accordance with established methodologies, we collected a comprehensive dataset of above- and below-ground functional traits for both plant species. These trait parameters are widely recognized in ecological research as key indicators of plants’ acquisition, utilization, and conservation strategies for critical resources41,53,54,55,56,57,58. Specifically, each stainless steel pot was fully submerged in water, and soil particles were removed under running tap water (water pressure: 0.2 MPa; duration: 5 min per sample). This process yielded intact plant specimens completely free of adhering soil particles.Randomly selected 4 intact leaves from the uppermost canopy, and collected 3 intact root segments representing the complete root system architecture8,31. Arranged samples on A4-sized acrylic trays with minimal overlap, scanned using a calibrated scanner (600 dpi, Epson 1680, Seiko Epson Corporation) following standardized calibration procedures, and analyzed using WinRHIZO software (Regent Instruments, Quebec, Canada) to determine leaf area (LA), average root diameter (RD), and total root length (RL)56. The scanned leaves and roots were dried to a constant mass in a 70 ℃ oven, then weighed, and these data were used to determine the specific leaf area (SLA) and specific root length (SRL)18,29,59.After these measurements, the residual parts of the plant material were divided into three parts: roots, stems, and leaves. Then place it in a 70 ℃ oven to dry until a constant weight was reached, and weigh it. We separately measured the total biomass of individual plants of A. philoxeroides and L. peploides. The leaf mass fraction (LMF) was calculated as the ratio of the leaf mass to the total mass32; The root mass fraction (RMF) was calculated as the ratio of the root mass to the total mass28.Collection and quantification of secondary metabolites in roots and leavesFollowing previous research methods60,61, randomly select 3–6 mature and intact leaves from harvested plants of weigh 1 g, and immediately grind them in a 10 ml sterile centrifuge tube. Add 5 ml of ethyl acetate for extraction and immerse for 72 h. Similarly, for collected belowground plant tissues and randomly select 2–5 intact root systems of weigh 1 g, and repeat the aforementioned steps. Preserve the obtained extracts in the dark at a low temperature (-20℃). Subsequently, filter the excess cellular debris using a 0.45 μm syringe filter, concentrate using a rotary evaporator (RE-52AA), dissolve in 1 ml of ethyl acetate, transfer to a GC-MS sample vial, and store in the dark at a low temperature (-40℃) until analysis62. The ethyl acetate extract of each sample was analysed by GC-MS (GCMS-QP 2020NX, SHIMADZU, Japan)63. The GC injector temperature was 250 °C. The oven temperature was maintained at 45 °C, then increased from 45 °C to 150 °C at 10 °C/min, and then increased to 250 °C at 15 °C/min for 10 min. The transfer line temperature was set to 250 °C. Helium was used as the carrier gas at a flow rate of 1 mL/min. The MS source was operated in electron impact (EI) mode at 70 eV. The MS was scanned from 45 to 450 m/z. For the GC–MS data of root exudates, we normalized the peak area of each compound with the sum of all peak area for allidentified metabolites for each sample. Relative peak area was then used to calculate compound-specific concentrations8,60,61,62,63.Determination of non-structural carbohydrates and C, N, P elementsThe total carbon and nitrogen concentrations in leaf and root plant tissues were analyzed using an organic elemental analyzer (Elementar, UNICUBE, Germany) via the dry combustion method. Total phosphorus content was measured using the molybdenum antimony anti-colorimetric method64. For the extraction and determination of soluble sugars and starch, 80% anhydrous ethanol and ethyl anthraquinone acetate reagents were utilized, respectively45. The sum of soluble sugar and starch content was considered the concentration of non-structural carbohydrates (NSC)65.Data analysisSince the data from each pot were not completely independent, prior to statistical analysis, we calculated the mean values of measurements for each plant species within individual pots. This approach ensures the validity of our statistical analysis.We employed two-way ANOVA to examine the effects of nitrogen addition, invasion degree, and their interaction on the total biomass, root traits, and leaf traits of A. philoxeroides and L. peploides. Means were compared using Duncan’s multiple range test. P-values ≤ 0.05 were considered statistically significant. Prior to analyses, we tested whether the assumptions of an ANOVA, homogeneity of variances and normally distributed residuals were achieved. The homogeneity of variances for all the studied parameters was evaluated by Levene’s test and the distribution of the residuals was assessed by Kolmogorov-Smirnov test. When necessary, logarithmic, reciprocal, or square root transformations were applied to meet assumptions.Based on their chemical properties, plant tissue metabolites were classified into six categories: phenols, alkaloids, amides, alkanes, organic acids, and terpenoids. Each value represents the total concentration of all compounds within a given category. Principal Component Analysis (PCA) was performed to visualize differences in exudate composition between the two plant species under varying nitrogen levels and invasion degree. Adonis analysis was used to test for significant differences between the nitrogen level and invasion degree. At the same time, two-way ANOVA was performed for each compound using the “dplyr” and “agricolae” packages in R 4.2.3 (R, 2022) to analyze the significant differences in the relative concentrations of the compounds, with nitrogen addition and invasion degree as independent variables.To examine the relationships between root/leaf functional traits and total biomass, we conducted Pearson correlation analyses separately for A. philoxeroides and L. peploides, assessing the associations between their respective root/leaf traits and total biomass. Furthermore, we quantified pairwise correlations between root and leaf traits under different nitrogen levels and invasion degree using Pearson’s correlation coefficients. This approach allowed us to investigate how these two species coordinate their root-leaf trait relationships in response to environmental changes.To further investigate the potential relationships among plant biomass, root traits, and leaf traits under different environmental factors, we employed partial least squares path modeling (PLS-PM) to evaluate the direct and indirect effects of invasion degree and nitrogen level on root traits, leaf traits, and total biomass in A. philoxeroides and L. peploides. The model was constructed using the “innerplot” function from the “plspm” package in R software (4.2.3). Hypothesized pathways were defined a priori based on ecological theory, with invasion degree and nitrogen levels as exogenous variables, and root traits, leaf traits, and total biomass as endogenous variables. All variables were standardized (mean = 0, SD = 1) to ensure comparability of path coefficients. Non-normality was addressed using the package’s robust weighting algorithm.All statistical analyses were performed using SPSS (SPSS Inc.) and R software (4.2.3), while figures were generated using Origin (Version 9.0, OriginLab Co.) and R software (4.2.3)63.

    Data availability

    The datasets generated and analyzed during the present study are accessible from the corresponding author upon reasonable request.
    ReferencesRen, G. Q. et al. Warming and elevated nitrogen deposition accelerate the invasion process of Solidago Canadensis L. Ecol. Process. 11, 3879 (2022).Article 

    Google Scholar 
    Sun, K. et al. Relative abundance of invasive plants more effectively explains the response of wetland communities to different invasion degrees than phylogenetic evenness. J. Plant. Ecol. 15, 625–638 (2022).Article 

    Google Scholar 
    Siebenkas, A., Schumacher, J. & Roscher, C. Phenotypic plasticity to light and nutrient availability alters functional trait ranking across eight perennial grassland species. AoB Plants. 7, plv029 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ni, M., Liu, Y., Chu, C. J., Xu, H. & Fang, S. Q. Fast seedling root growth leads to competitive superiority of invasive plants. Biol. Invasions. 20, 1821–1832 (2018).Article 

    Google Scholar 
    Chen, L. H. et al. Effects of simulated nitrogen deposition on the ecophysiological responses of Populus beijingensis and P. cathayana under intra- and interspecific competition. Plant. Soil. 481, 127–146 (2022).Article 
    CAS 

    Google Scholar 
    Li, L., Ding, M. M. & Jeppesen, E. Variation in growth, reproduction, and resource allocation in an aquatic plant, Vallisneria spinulosa: the influence of amplitude and frequency of water level fluctuations. Aquat. Sci. 82, 983 (2020).Article 

    Google Scholar 
    Read, Q. D., Henning, J. A., Classen, A. T. & Sanders, N. J. Aboveground resilience to species loss but belowground resistance to nitrogen addition in a montane plant community. J. Plant. Ecol. 11, 351–363 (2018).Article 

    Google Scholar 
    Li, D. X. et al. The impacts of different nitrogen supply on root traits, root exudates, and soil enzyme activities of exotic and native plant communities. Plant. Soil. 508, 209–226 (2024).Article 

    Google Scholar 
    Wang, T., Hu, J. T., Miao, L. L., Yu, D. & Liu, C. H. The invasive stoloniferous clonal plant Alternanthera Philoxeroides outperforms its co-occurring non-invasive functional counterparts in heterogeneous soil environments – invasion implications. Sci. Rep. 6, 38036 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, P. et al. Contrasting coordination of non-structural carbohydrates with leaf and root economic strategies of alpine coniferous forests. New. Phytol. 243, 580–590 (2024).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sun, J. K. et al. Advantages of growth and competitive ability of the invasive plant Solanum rostratum over two co-occurring natives and the effects of nitrogen levels and forms. Front. Plant. Sci. 14, 1169317 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, J. L., Huang, W. & Ding, J. Q. Phenotypic plasticity in resource allocation to sexual trait of alligatorweed in wetland and terrestrial habitats. Sci. Total Environ. 757, 143819 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Si, C. C. et al. Different degrees of plant invasion significantly affect the richness of the soil fungal community. PLoS One. 8, e85490 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, H. Y. et al. Invasion by the weed Conyza Canadensis alters soil nutrient supply and shifts microbiota structure. Soil. Biol. Biochem. 143, 107739 (2020).Article 
    CAS 

    Google Scholar 
    Xing, L. J. et al. Comparison between the exotic Coreopsis grandiflora and native Dendranthema indicum across variable nitrogen deposition conditions. Acta Physiol. Plant. 44, 349 (2022).Article 

    Google Scholar 
    Guo, X. et al. Nitrogen deposition effects on invasive and native plant competition: implications for future invasions. Ecotoxicol. Environ. Saf. 259, 115029 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Xu, F. W. et al. Linking leaf traits to the Temporal stability of above- and belowground productivity under global change and land use scenarios in a semi-arid grassland of inner Mongolia. Sci. Total Environ. 818, 151858 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Gong, J. R. et al. N addition rebalances the carbon and nitrogen metabolisms of Leymus chinensis through leaf N investment. Plant. Physiol. Biochem. 185, 221–232 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wu, H. et al. Nitrogen enrichment alters the resistance of a noninvasive alien plant species to Alternanthera Philoxeroides invasion. Front. Ecol. Evol. 11, 1215191 (2023).Article 

    Google Scholar 
    Wang, B. et al. Nitrogen addition alters photosynthetic carbon fixation, allocation of photoassimilates, and carbon partitioning of Leymus chinensis in a temperate grassland of inner Mongolia. Agric. Meteorol. 279, 107743 (2019).Article 

    Google Scholar 
    Zhang, Q. Z. et al. Nitrogen addition and drought affect nitrogen uptake patterns and biomass production of four urban greening tree species in North China. Sci. Total Environ. 893, 164893 (2023).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhao, Z. W. et al. Effects of nitrogen addition on plant-soil-microbe stoichiometry characteristics of different functional group species in Bothriochloa ischemum community. Soil. Ecol. Lett. 4, 362–375 (2021).Article 

    Google Scholar 
    Jamieson, M. A., Seastedt, T. R. & Bowers, M. D. Nitrogen enrichment differentially affects above- and belowground plant defense. Am. J. Bot. 99, 1630–1637 (2012).Article 
    PubMed 

    Google Scholar 
    Wang, A. O. et al. Nitrogen addition increases intraspecific competition in the invasive wetland plant Alternanthera philoxeroides, but not in its native congener Alternanthera sessilis. Plant. Species Biol. 30, 176–183 (2014).Article 

    Google Scholar 
    Wei, X. W. et al. Improved utilization of nitrate nitrogen through within-leaf nitrogen allocation trade-offs in Leymus chinensis. Front. Plant. Sci. 13, 870681 (2022).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hu, Y. K. et al. Is there coordination of leaf and fine root traits at local scales? A test in temperate forest swamps. Ecol. Evol. 9, 8714–8723 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hu, Y. K., Pan, X., Liu, X. Y., Fu, Z. X. & Zhang, M. Y. Above- and belowground plant functional composition show similar changes during temperate forest swamp succession. Front. Plant. Sci. 12, 658883 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wang, Q. et al. Trait value and phenotypic integration contribute to the response of exotic Rhus typhina to heterogeneous nitrogen deposition: a comparison with native Rhus chinensis. Sci. Total Environ. 844, 157199 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    de la Riva, E. G. et al. Root traits across environmental gradients in mediterranean Woody communities: are they aligned along the root economics spectrum? Plant. Soil. 424, 35–48 (2017).Article 

    Google Scholar 
    Du, L. S., Liu, H. Y., Guan, W. B., Li, J. M. & Li, J. S. Drought affects the coordination of belowground and aboveground resource-related traits in Solidago Canadensis in China. Ecol. Evol. 9, 9948–9960 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jo, I., Fridley, J. D. & Frank, D. A. Linking above- and belowground resource use strategies for native and invasive species of temperate deciduous forests. Biol. Invasions. 17, 1545–1554 (2014).Article 

    Google Scholar 
    Ren, G. Q. et al. The enhancement of root biomass increases the competitiveness of an invasive plant against a co-occurring native plant under elevated nitrogen deposition. Flora 261, 151486 (2019).Article 

    Google Scholar 
    Xu, K. et al. Nitrogen deposition further increases Ambrosia trifida root exudate invasiveness under global warming. Environ. Monit. Assess. 195, 759 (2023).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wang, C. Y., Zhou, J. W., Liu, J. & Jiang, K. Differences in functional traits between invasive and native Amaranthus species under different forms of N deposition. Sci. Nat. 104, 927 (2017).Article 

    Google Scholar 
    Witzell, J. & Shevtsova, A. Nitrogen-induced changes in phenolics of Vaccinium myrtillus–implications for interaction with a parasitic fungus. J. Chem. Ecol. 30, 1937–1956 (2004).Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhou, J. L. et al. Nitrogen influence to the independent invasion and the co-invasion of Solidago Canadensis and Conyza Canadensis via intensified allelopathy. Sustainability 14, 11970 (2022).Article 
    CAS 

    Google Scholar 
    Liese, R., Alings, K. & Meier, I. C. Root branching is a leading root trait of the plant economics spectrum in temperate trees. Front. Plant. Sci. 8, 315 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Martin, A. R. et al. Integrating nitrogen fixing structures into above- and belowground functional trait spectra in soy (Glycine max). Plant. Soil. 440, 53–69 (2019).Article 
    CAS 

    Google Scholar 
    van Geem, M. et al. The importance of aboveground-belowground interactions on the evolution and maintenance of variation in plant defense traits. Front. Plant. Sci. 4, 431 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Shi, M. Z. et al. Plant volatile compounds of the invasive alligatorweed, Alternanthera Philoxeroides (Mart.) Griseb, infested by Agasicles hygrophila Selman and Vogt (Coleoptera: Chrysomelidae). Life (Basel). 12, 1257 (2022).CAS 
    PubMed 

    Google Scholar 
    Weemstra, M. et al. Weak phylogenetic and habitat effects on root trait variation of 218 Neotropical tree species. Front. Glob Change. 6, 7127 (2023).Article 

    Google Scholar 
    Roiloa, S. R., Rodríguez-Echeverría, S., Freitas, H. & Retuerto, R. Developmentally-programmed division of labour in the clonal invader Carpobrotus Edulis. Biol. Invasions. 15, 1895–1905 (2013).Article 

    Google Scholar 
    Huang, J., Wang, X. M., Zheng, M. H. & Mo, J. M. 13-year nitrogen addition increases nonstructural carbon pools in subtropical forest trees in Southern China. Ecol. Manag. 481, 118748 (2021).Article 

    Google Scholar 
    Du, Y., Lu, R. L. & Xia, J. Y. Impacts of global environmental change drivers on non-structural carbohydrates in terrestrial plants. Funct. Ecol. 34, 1525–1536 (2020).Article 

    Google Scholar 
    Yin, J. L. et al. Carbohydrate, phytohormone, and associated transcriptome changes during storage root formation in alligatorweed (Alternanthera philoxeroides). Weed Sci. 68, 382–395 (2020).Article 

    Google Scholar 
    Geng, Y. P. et al. Plasticity and ontogenetic drift of biomass allocation in response to above- and below-ground resource availabilities in perennial herbs: a case study of Alternanthera Philoxeroides. Ecol. Res. 22, 255–260 (2006).Article 

    Google Scholar 
    Smith, M. S., Fridley, J. D., Goebel, M. & Bauerle, T. L. Links between belowground and aboveground resource-related traits reveal species growth strategies that promote invasive advantages. PLoS One. 9, e104189 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zhang, Y. J. et al. Vertical distribution patterns of community biomass, carbon and nitrogen content in grasslands on the Eastern Qinghai–Tibet plateau. Ecol. Indic. 154, 110726 (2023).Article 
    CAS 

    Google Scholar 
    Xian, L. et al. Which has a greater impact on plant functional traits: plant source or environment? Plants (Basel). 13, 903 (2024).CAS 
    PubMed 

    Google Scholar 
    Fan, R., Hua, J. G., Huang, Y. L., Lin, J. Y. & Ji, W. L. What role do dauciform roots play? Responses of Carex filispica to trampling in alpine meadows based on functional traits. Ecol. Evol. 13, e9875 (2023).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Marques, E. et al. The impact of domestication on aboveground and belowground trait responses to nitrogen fertilization in wild and cultivated genotypes of Chickpea (Cicer sp). Front. Genet. 11, 576338 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Huang, X. L. et al. Root morphological and structural comparisons of introduced and native aquatic plant species in multiple substrates. Aquat. Ecol. 52, 65–76 (2017).Article 

    Google Scholar 
    Fry, E. L. et al. Soil multifunctionality and drought resistance are determined by plant structural traits in restoring grassland. Ecology 99, 2260–2271 (2018).Article 
    PubMed 

    Google Scholar 
    Weemstra, M. et al. Towards a multidimensional root trait framework: a tree root review. New. Phytol. 211, 1159–1169 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wu, D. et al. Plant phenols contents and their changes with nitrogen availability in peatlands of Northeastern China. J. Plant. Ecol. 13, 713–721 (2020).Article 
    CAS 

    Google Scholar 
    Lin, G. G., Gao, M. X., Zeng, D. H. & Fang, Y. T. Aboveground conservation acts in synergy with belowground uptake to alleviate phosphorus deficiency caused by nitrogen addition in a larch plantation. Ecol. Manag. 473, 118309 (2020).Article 

    Google Scholar 
    Lv, T. et al. Invasive submerged plant has a stronger inhibitory effect on epiphytic algae than native plant. Biol. Invasions. 26, 1001–1014 (2023).Article 

    Google Scholar 
    Gervais-Bergeron, B., Chagnon, P. L. & Labrecque, M. Willow aboveground and belowground traits can predict phytoremediation services. Plants (Basel). 10, 1824 (2021).CAS 
    PubMed 

    Google Scholar 
    Bakker, L. M., Mommer, L. & van Ruijven, J. Using root traits to understand Temporal changes in biodiversity effects in grassland mixtures. Oikos 128, 208–220 (2018).Article 

    Google Scholar 
    Bi, J. W. et al. Divergent geographic variation in above- versus below‐ground secondary metabolites of Reynoutria Japonica. J. Ecol. 112, 514–527 (2024).Article 

    Google Scholar 
    Yu, H. W. et al. Greater chemical signaling in root exudates enhances soil mutualistic associations in invasive plants compared to natives. New. Phytol. 236, 1140–1153 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Vivanco, J. M., Bais, H. P., Stermitz, F. R., Thelen, G. C. & Callaway, R. M. Biogeographical variation in community response to root allelochemistry: novel weapons and exotic invasion. Ecol. Lett. 7, 285–292 (2004).Article 

    Google Scholar 
    Lv, T. et al. Invasion of water hyacinth and water lettuce inhibits the abundance of epiphytic algae. Divers. Distrib. 28, 1650–1662 (2022).Article 

    Google Scholar 
    Chen, C., Xing, F., Li, Z. & Zhang, R. H. Nitrogen addition changes the allelopathic effects of the root leachate from the invasive weed Stellera Chamaejasme L. on a dominant grass in the Songnen grassland. J. Plant. Biol. 66, 243–255 (2023).Article 
    CAS 

    Google Scholar 
    Zhou, L. F. et al. Latitudinal and longitudinal trends of seed traits indicate adaptive strategies of an invasive plant. Front. Plant. Sci. 12, 657813 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Download referencesAcknowledgementsThe authors gratefully acknowledge funding support from the Fundamental Research Funds for the Central Universities (2042020kf1025).Author informationAuthors and AffiliationsDepartment of Ecology, College of Life Science, The National Field Station of Freshwater Ecosystem of Liangzi Lake, Wuhan University, 299 Bayi Rd., Wuhan, ChinaDexiang Li, Tian Lv, Yang Li, Haihao Yu, Dan Yu & Chunhua LiuAuthorsDexiang LiView author publicationsSearch author on:PubMed Google ScholarTian LvView author publicationsSearch author on:PubMed Google ScholarYang LiView author publicationsSearch author on:PubMed Google ScholarHaihao YuView author publicationsSearch author on:PubMed Google ScholarDan YuView author publicationsSearch author on:PubMed Google ScholarChunhua LiuView author publicationsSearch author on:PubMed Google ScholarContributionsDexiang Li led the research design. Dexiang Li, Tian Lv and Yang Li collected raw data set and integrated. Dexiang Li conducted data management, led a statistical analysis and led the writing of the first manuscript. Haihao Yu, Dan Yu, and Chunhua Liu commented on previous versions of the manuscript. All authors interpreted the results and significantly contributed to improve the manuscript. All authors read and approved the final manuscript.Corresponding authorCorrespondence to
    Chunhua Liu.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Statement
    The collection of plant materials (Alternanthera philoxeroides and Ludwigia peploides) and experimental research conducted in this study comply with relevant institutional, national, and international guidelines and regulations. We confirm that this study adheres to all applicable institutional, national, and international standards and legislation.

    Plant guidelines
    The plant materials used in this study (Alternanthera philoxeroides and Ludwigia peploides) were formally identified by Professor Dan Yu from the Department of Aquatic Ecology, College of Life Sciences, Wuhan University, to ensure species accuracy. Professor Dan Yu (email: [email protected]), a professor in the Department of Aquatic Ecology, College of Life Sciences, Wuhan University, specializes in plant taxonomy and identification. Prior to experimentation, both plant species used in this study were authenticated by Prof. Yu to ensure taxonomic accuracy. Additionally, voucher specimens of both A. philoxeroides and L. peploides have been deposited at the Liangzi Lake National Field Station for Freshwater Ecosystem Research, Wuhan University. The voucher specimens have been deposited with the following accession numbers, A. philoxeroides: Collection No. WHU-AP-20220401-001; L. peploides: Collection No. WHU-LP-20220401-001. This research obtained the necessary collection permits for Alternanthera philoxeroides and Ludwigia peploides from the Liangzi Lake National Field Station for Freshwater Ecosystem Research, Wuhan University.

    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 articleLi, D., Lv, T., Li, Y. et al. Functional trait variations of the invasive plant Alternanthera Philoxeroides and the native plant Ludwigia peploides under nitrogen addition.
    Sci Rep 15, 43799 (2025). https://doi.org/10.1038/s41598-025-27758-4Download citationReceived: 17 June 2025Accepted: 05 November 2025Published: 15 December 2025Version of record: 15 December 2025DOI: https://doi.org/10.1038/s41598-025-27758-4Share 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
    KeywordsFunctional traitsNitrogen depositionInvasion processCoordination More

  • in

    New SAR11 isolate genomes and global marine metagenomes resolve ecologically relevant units within the Pelagibacterales

    AbstractThe bacterial order Pelagibacterales (SAR11) is widely distributed across the global surface ocean, where its activities are integral to the marine carbon cycle. High-quality genomes from isolates that can be propagated and phenotyped are needed to unify perspectives on the ecology and evolution of this complex group. Here, we increase the number of complete SAR11 isolate genomes threefold by describing 81 new SAR11 strains from coastal and offshore surface seawater of the tropical Pacific Ocean. Our analyses of the genomes and their spatiotemporal distributions support the existence of 29 monophyletic, discrete Pelagibacterales ecotypes that we define as genera. The spatiotemporal distributions of genomes within genera were correlated at fine scales with variation in ecologically-relevant gene content, supporting generic assignments and providing indications of speciation. We provide a hierarchical system of classification for SAR11 populations that is meaningfully correlated with evolution and ecology, providing a valid and utilitarian systematic nomenclature for this clade.

    Data availability

    We have deposited the assembled sequence data for newly sequenced genomes, including raw sequencing reads, under NCBI BioProject ID PRJNA1170004. Ribosomal RNA gene sequence data were published previously under NCBI BioProject ID PRJNA673898. The Supplementary Data includes all remaining data, including accession numbers for all previously sequenced genomes and metagenomes. The URL https://seqco.de/r:r4auejub serves as the SeqCode registry for all taxon names defined in this study.
    Code availability

    All custom R, BASH, and Python scripts used for data analyses in this study are publicly available on GitHub at https://github.com/kcfreel/SAR11-genomes-from-the-tropical-Pacific and archived with Zenodo (https://doi.org/10.5281/zenodo.17614008). Additionally, a fully reproducible bioinformatics workflow for the analysis of SAR11 genomes is available at https://merenlab.org/data/sar11-phylogenomics/, enabling the reproduction of our phylogenomic tree and its extension with new genomes.
    ReferencesGrote, J. et al. Streamlining and core genome conservation among highly divergent members of the SAR11 clade. mBio 3, e00252–12 (2012).Morris, R. M. et al. SAR11 clade dominates ocean surface bacterioplankton communities. Nature 420, 806–810 (2002).
    Google Scholar 
    Carlson, C. A. et al. Seasonal dynamics of SAR11 populations in the euphotic and mesopelagic zones of the northwestern Sargasso Sea. ISME J. 3, 283–295 (2009).
    Google Scholar 
    Schattenhofer, M. et al. Latitudinal distribution of prokaryotic picoplankton populations in the Atlantic Ocean. Environ. Microbiol. 11, 2078–2093 (2009).
    Google Scholar 
    Eiler, A., Hayakawa, D. H., Church, M. J., Karl, D. M. & Rappé, M. S. Dynamics of the SAR11 bacterioplankton lineage in relation to environmental conditions in the oligotrophic North Pacific subtropical gyre. Environ. Microbiol. 11, 2291–2300 (2009).
    Google Scholar 
    Becker, J. W., Hogle, S. L., Rosendo, K. & Chisholm, S. W. Co-culture and biogeography of Prochlorococcus and SAR11. ISME J. 13, 1506–1519 (2019).
    Google Scholar 
    Eren, A. M. et al. Oligotyping: differentiating between closely related microbial taxa using 16S rRNA gene data. Methods Ecol. Evol. 4, 1111–1119 (2013).
    Google Scholar 
    Delmont, T. O. et al. Single-amino acid variants reveal evolutionary processes that shape the biogeography of a global SAR11 subclade. Elife 8, e46497 (2019).Tucker, S. J. et al. Spatial and temporal dynamics of SAR11 marine bacteria across a nearshore to offshore transect in the tropical Pacific Ocean. PeerJ 9, e12274 (2021).
    Google Scholar 
    Haro-Moreno, J. M. et al. Ecogenomics of the SAR11 clade. Environ. Microbiol. 22, 1748–1763 (2020).
    Google Scholar 
    Tucker, S. J., Freel, K. C., Eren, A. M. & Rappé, M. S. Habitat-specificity in SAR11 is associated with a few genes under high selection. ISME J. 19, wraf216 https://doi.org/10.1093/ismejo/wraf216 (2025).Giovannoni, S. J., Britschgi, T. B., Moyer, C. L. & Field, K. G. Genetic diversity in Sargasso Sea bacterioplankton. Nature 345, 60–63 (1990).
    Google Scholar 
    Hug, L. A. et al. A new view of the tree of life. Nat. Microbiol. 1, 16048 (2016).
    Google Scholar 
    Paoli, L. et al. Biosynthetic potential of the global ocean microbiome. Nature 607, 111–118 (2022).
    Google Scholar 
    Tsementzi, D. et al. SAR11 bacteria linked to ocean anoxia and nitrogen loss. Nature 536, 179–183 (2016).
    Google Scholar 
    Kiefl, E. et al. Structure-informed microbial population genetics elucidates selective pressures that shape protein evolution. Sci. Adv. 9, eabq4632 (2023).
    Google Scholar 
    López-Pérez, M., Haro-Moreno, J. M., Coutinho, F. H., Martinez-Garcia, M. & Rodriguez-Valera, F. The evolutionary success of the marine bacterium SAR11 analyzed through a metagenomic perspective. mSystems 5, e00605–20 (2020).Delmont, T. O. et al. Nitrogen-fixing populations of Planctomycetes and Proteobacteria are abundant in surface ocean metagenomes. Nat. Microbiol. 3, 804–813 (2018).
    Google Scholar 
    Tully, B. J., Graham, E. D. & Heidelberg, J. F. The reconstruction of 2631 draft metagenome-assembled genomes from the global oceans. Sci. Data 5, 170203 (2018).
    Google Scholar 
    Chang, T., Gavelis, G. S., Brown, J. M. & Stepanauskas, R. Genomic representativeness and chimerism in large collections of SAGs and MAGs of marine prokaryoplankton. Microbiome 12, 126 (2024).
    Google Scholar 
    Pachiadaki, M. G. et al. Charting the complexity of the marine microbiome through single-cell genomics. Cell 179, 1623–1635.e11 (2019).
    Google Scholar 
    Thrash, J. C. et al. Phylogenomic evidence for a common ancestor of mitochondria and the SAR11 clade. Sci. Rep. 1, 13 (2011).
    Google Scholar 
    Muñoz-Gómez, S. A. et al. An updated phylogeny of the Alphaproteobacteria reveals that the parasitic Rickettsiales and Holosporales have independent origins. Elife 8, e42535 (2019).Vergin, K. L. et al. High intraspecific recombination rate in a native population of Candidatus pelagibacter ubique (SAR11). Environ. Microbiol. 9, 2430–2440 (2007).
    Google Scholar 
    Wilhelm, L. J., Tripp, H. J., Givan, S. A., Smith, D. P. & Giovannoni, S. J. Natural variation in SAR11 marine bacterioplankton genomes inferred from metagenomic data. Biol. Direct 2, 27 (2007).
    Google Scholar 
    Carini, P., Steindler, L., Beszteri, S. & Giovannoni, S. J. Nutrient requirements for growth of the extreme oligotroph ‘Candidatus Pelagibacter ubique’ HTCC1062 on a defined medium. ISME J. 7, 592–602 (2013).
    Google Scholar 
    Sun, J. et al. The abundant marine bacterium Pelagibacter simultaneously catabolizes dimethylsulfoniopropionate to the gases dimethyl sulfide and methanethiol. Nat. Microbiol. 1, 16065 (2016).
    Google Scholar 
    Tripp, H. J. et al. SAR11 marine bacteria require exogenous reduced sulphur for growth. Nature 452, 741–744 (2008).
    Google Scholar 
    Rappé, M. S., Connon, S. A., Vergin, K. L. & Giovannoni, S. J. Cultivation of the ubiquitous SAR11 marine bacterioplankton clade. Nature 418, 630–633 (2002).
    Google Scholar 
    Giovannoni, S. J. et al. Genome streamlining in a cosmopolitan oceanic bacterium. Science 309, 1242–1245 (2005).
    Google Scholar 
    Schwalbach, M. S., Tripp, H. J., Steindler, L., Smith, D. P. & Giovannoni, S. J. The presence of the glycolysis operon in SAR11 genomes is positively correlated with ocean productivity. Environ. Microbiol. 12, 490–500 (2010).
    Google Scholar 
    Sun, J. et al. One-carbon metabolism in SAR11 pelagic marine bacteria. PLoS ONE 6, e23973 (2011).
    Google Scholar 
    Giovannoni, S. J. SAR11 Bacteria: The Most Abundant Plankton in the Oceans. Ann. Rev. Mar. Sci. 9, 231–255 (2017).
    Google Scholar 
    Giovannoni, S. J., Cameron Thrash, J. & Temperton, B. Implications of streamlining theory for microbial ecology. ISME J. 8, 1553–1565 (2014).
    Google Scholar 
    Viklund, J., Ettema, T. J. G. & Andersson, S. G. E. Independent genome reduction and phylogenetic reclassification of the oceanic SAR11 clade. Mol. Biol. Evol. 29, 599–615 (2012).
    Google Scholar 
    Carini, P. et al. Discovery of a SAR11 growth requirement for thiamin’s pyrimidine precursor and its distribution in the Sargasso Sea. ISME J. 8, 1727–1738 (2014).
    Google Scholar 
    Brandon, M. L. High-Throughput Isolation of Pelagic Marine Bacteria from the Coastal Subtropical Pacific Ocean Master’s thesis (University of Hawaiʻi at Mānoa, Department of Oceanography, 2006).Vergin, K. L. et al. High-resolution SAR11 ecotype dynamics at the Bermuda Atlantic Time-series study site by phylogenetic placement of pyrosequences. ISME J. 7, 1322–1332 (2013).
    Google Scholar 
    Thrash, J. C. et al. Single-cell enabled comparative genomics of a deep ocean SAR11 bathytype. ISME J. 8, 1440–1451 (2014).
    Google Scholar 
    Wang, Z. & Wu, M. A phylum-level bacterial phylogenetic marker database. Mol. Biol. Evol. 30, 1258–1262 (2013).
    Google Scholar 
    Suzuki, M. T., Béjà, O., Taylor, L. T. & Delong, E. F. Phylogenetic analysis of ribosomal RNA operons from uncultivated coastal marine bacterioplankton. Environ. Microbiol. 3, 323–331 (2001).
    Google Scholar 
    Getz, E. W. et al. The AEGEAN-169 clade of bacterioplankton is synonymous with SAR11 subclade V (HIMB59) and metabolically distinct. mSystems 8, e0017923 (2023).
    Google Scholar 
    Viklund, J., Martijn, J., Ettema, T. J. G. & Andersson, S. G. E. Comparative and phylogenomic evidence that the alphaproteobacterium HIMB59 is not a member of the oceanic SAR11 clade. PLoS ONE 8, e78858 (2013).
    Google Scholar 
    Martijn, J., Vosseberg, J., Guy, L., Offre, P. & Ettema, T. J. G. Deep mitochondrial origin outside the sampled alphaproteobacteria. Nature 557, 101–105 (2018).
    Google Scholar 
    Muñoz-Gómez, S. A. et al. Site-and-branch-heterogeneous analyses of an expanded dataset favour mitochondria as sister to known Alphaproteobacteria. Nat. Ecol. Evol. 6, 253–262 (2022).
    Google Scholar 
    Evans, J. T. & Denef, V. J. To dereplicate or not to dereplicate? mSphere 5, e00971–19 (2020).Zhao, J. et al. Promiscuous and genome-wide recombination underlies the sequence-discrete species of the SAR11 lineage in the deep ocean. ISME J. 19, wraf072 (2025).Hellweger, F. L., van Sebille, E. & Fredrick, N. D. Biogeographic patterns in ocean microbes emerge in a neutral agent-based model. Science 345, 1346–1349 (2014).
    Google Scholar 
    Villarreal-Chiu, J. F., Quinn, J. P. & McGrath, J. W. The genes and enzymes of phosphonate metabolism by bacteria, and their distribution in the marine environment. Front. Microbiol. 3, 19 (2012).
    Google Scholar 
    Carini, P., White, A. E., Campbell, E. O. & Giovannoni, S. J. Methane production by phosphate-starved SAR11 chemoheterotrophic marine bacteria. Nat. Commun. 5, 4346 (2014).
    Google Scholar 
    Sosa, O. A., Repeta, D. J., DeLong, E. F., Ashkezari, M. D. & Karl, D. M. Phosphate-limited ocean regions select for bacterial populations enriched in the carbon-phosphorus lyase pathway for phosphonate degradation. Environ. Microbiol. 21, 2402–2414 (2019).
    Google Scholar 
    Acker, M. et al. Phosphonate production by marine microbes: exploring new sources and potential function. Proc. Natl Acad. Sci. USA. 119, e2113386119 (2022).
    Google Scholar 
    Zhao, X. et al. Three-dimensional structure of the ultraoligotrophic marine bacterium ‘Candidatus Pelagibacter ubique’. Appl. Environ. Microbiol. 83, e02807–16 (2017).Craig, L. & Li, J. Type IV pili: paradoxes in form and function. Curr. Opin. Struct. Biol. 18, 267–277 (2008).
    Google Scholar 
    Braakman, R. et al. Global niche partitioning of purine and pyrimidine cross-feeding among ocean microbes. Sci. Adv. 11, eadp1949 (2025).Newton, R. J., Jones, S. E., Eiler, A., McMahon, K. D. & Bertilsson, S. A guide to the natural history of freshwater lake bacteria. Microbiol. Mol. Biol. Rev. 75, 14–49 (2011).
    Google Scholar 
    Brown, M. V. et al. Global biogeography of SAR11 marine bacteria. Mol. Syst. Biol. 8, 595 (2012).
    Google Scholar 
    Davies, T. J. Evolutionary ecology: when relatives cannot live together. Curr. Biol. 16, R645–R647 (2006).
    Google Scholar 
    Ramfelt, O., Freel, K. C., Tucker, S. J., Nigro, O. D. & Rappé, M. S. Isolate-anchored comparisons reveal evolutionary and functional differentiation across SAR86 marine bacteria. ISME J. 18, wrae227 (2024).Tucker, S. J. et al. Seasonal and spatial transitions in phytoplankton assemblages spanning estuarine to open ocean waters of the tropical Pacific. Limnol. Oceanogr. 70, 1693–1708 (2025).
    Google Scholar 
    Ramfelt, O., Tucker, S. J., Freel, K. C., Eren, A. M. & Rappe, M. S. Magnimaribacterales marine bacteria genetically partition across the nearshore to open-ocean in the tropical Pacific Ocean. Preprint at https://doi.org/10.1101/2025.06.17.660167 (2025).Jain, C., Rodriguez-R, L. M., Phillippy, A. M., Konstantinidis, K. T. & Aluru, S. High throughput ANI analysis of 90K prokaryotic genomes reveals clear species boundaries. Nat. Commun. 9, 5114 (2018).
    Google Scholar 
    Olm, M. R. et al. Consistent metagenome-derived metrics verify and delineate bacterial species boundaries. mSystems 5, e00731–19 (2020).Parks, D. H. et al. GTDB: an ongoing census of bacterial and archaeal diversity through a phylogenetically consistent, rank normalized and complete genome-based taxonomy. Nucleic Acids Res. 50, D785–D794 (2022).
    Google Scholar 
    Waite, D. W. et al. Proposal to reclassify the proteobacterial classes Deltaproteobacteria and Oligoflexia, and the phylum Thermodesulfobacteria into four phyla reflecting major functional capabilities. Int. J. Syst. Evol. Microbiol. 70, 5972–6016 (2020).
    Google Scholar 
    Sanford, R. A., Lloyd, K. G., Konstantinidis, K. T. & Löffler, F. E. Microbial taxonomy run amok. Trends Microbiol. 29, 394–404 (2021).
    Google Scholar 
    Monaghan, E. A., Freel, K. C. & Rappé, M. S. Isolation of SAR11 Marine Bacteria from Cryopreserved Seawater. mSystems 5, e00954–20 (2020).Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).
    Google Scholar 
    Callahan, B. J. et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).
    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).
    Google Scholar 
    Bankevich, A. et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 19, 455–477 (2012).
    Google Scholar 
    Wick, R. R., Judd, L. M., Gorrie, C. L. & Holt, K. E. Unicycler: resolving bacterial genome assemblies from short and long sequencing reads. PLoS Comput. Biol. 13, e1005595 (2017).
    Google Scholar 
    Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).
    Google Scholar 
    Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).
    Google Scholar 
    Eren, A. M. et al. Anvi’o: an advanced analysis and visualization platform for ‘omics data. PeerJ 3, e1319 (2015).
    Google Scholar 
    Eren, A. M. et al. Community-led, integrated, reproducible multi-omics with anvi’o. Nat. Microbiol. 6, 3–6 (2021).
    Google Scholar 
    Robinson, J. T. et al. Integrative genomics viewer. Nat. Biotechnol. 29, 24–26 (2011).
    Google Scholar 
    Milne, I. et al. Using tablet for visual exploration of second-generation sequencing data. Brief. Bioinform. 14, 193–202 (2013).
    Google Scholar 
    Parks, D. H., Imelfort, M., Skennerton, C. T., Hugenholtz, P. & Tyson, G. W. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 25, 1043–1055 (2015).
    Google Scholar 
    Chaumeil, P.-A., Mussig, A. J., Hugenholtz, P. & Parks, D. H. GTDB-Tk: a toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics 36, 1925–1927 (2019).
    Google Scholar 
    Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).
    Google Scholar 
    Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).
    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, 1530–1534 (2020).
    Google Scholar 
    Kalyaanamoorthy, S., Minh, B. Q., Wong, T. K. F., von Haeseler, A. & Jermiin, L. S. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat. Methods 14, 587–589 (2017).
    Google Scholar 
    Revell, L. J. phytools 2.0: an updated R ecosystem for phylogenetic comparative methods (and other things). PeerJ 12, e16505 (2024).
    Google Scholar 
    R Development Core Team, R. R: A Language and Environment for Statistical Computing https://doi.org/10.1007/978-3-540-74686-7 (2011).Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—Approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).
    Google Scholar 
    Parks, D. H. et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat. Biotechnol. 36, 996–1004 (2018).
    Google Scholar 
    Winter, K. B. et al. Collaborative Research To Inform Adaptive Comanagement: A Framework For The Heʻeia National Estuarine Research Reserve. Ecol. Soc. 25, 15 (2020).Kūlana Noiʻi Working Group. Kūlana Noiʻi version 2. [online] URL: https://seagrant.soest.hawaii.edu/wp-content/uploads/2021/09/Kulana-Noii-2.0_LowRes.pdf (Hawaiʻi Sea Grant, 2021).Sunagawa, S. et al. Ocean plankton. Structure and function of the global ocean microbiome. Science 348, 1261359 (2015).
    Google Scholar 
    Mende, D. R. et al. Environmental drivers of a microbial genomic transition zone in the ocean’s interior. Nat. Microbiol. 2, 1367–1373 (2017).
    Google Scholar 
    Biller, S. J. et al. Marine microbial metagenomes sampled across space and time. Sci. Data 5, 180176 (2018).
    Google Scholar 
    Kudo, T. et al. Seasonal changes in the abundance of bacterial genes related to dimethylsulfoniopropionate catabolism in seawater from Ofunato Bay revealed by metagenomic analysis. Gene 665, 174–184 (2018).
    Google Scholar 
    Yoshitake, K. et al. Development of a time-series shotgun metagenomics database for monitoring microbial communities at the Pacific coast of Japan. Sci. Rep. 11, 12222 (2021).
    Google Scholar 
    Mueller, R. S. et al. Metagenome sequencing of a coastal marine microbial community from Monterey Bay, California. Genome Announc. 3, e00341–15 (2015).Kopf, A. et al. The Ocean Sampling Day Consortium. Gigascience 4, 27 (2015).
    Google Scholar 
    Shaiber, A. et al. Functional and genetic markers of niche partitioning among enigmatic members of the human oral microbiome. Genome Biol. 21, 292 (2020).
    Google Scholar 
    Köster, J. & Rahmann, S. Building and documenting workflows with Python-based snakemake. GCB 49, 56 (2012).
    Google Scholar 
    Eren, A. M., Vineis, J. H., Morrison, H. G. & Sogin, M. L. A filtering method to generate high-quality short reads using Illumina paired-end technology. PLoS ONE 8, e66643 (2013).
    Google Scholar 
    Utter, D. R. et al. Metapangenomics of the oral microbiome provides insights into habitat adaptation and cultivar diversity. Genome Biol. 21, 293 (2020).
    Google Scholar 
    Community Ecology Package [R package vegan version 2.7-1]. Comprehensive R Archive Network (CRAN) https://cran.r-project.org/web/packages/vegan/index.html (2025).Delmont, T. O. & Eren, A. M. Linking pangenomes and metagenomes: the Prochlorococcus metapangenome. PeerJ https://doi.org/10.7717/peerj.4320 (2018).Download referencesAcknowledgementsWe thank Kumu Hula, Kawaikapuokalani Hewett and Aimee Sato for their generous guidance in using appropriate ʻŌlelo Hawaiʻi (Hawaiian Language) words to create species names for isolates cultivated on Moku o Loʻe. We also thank K. Luttrell for help with Latin grammar, R. Malmstrom and N. Nath for sequencing the genomes of isolates HIMB109 and HIMB123, R. Ouye for assistance with HTC experiments, O. Ramfelt for bioinformatic support, and C. Foley for her generous help with creating the map used in Supplementary Fig. 1b. We also thank F. Trigodet for their help with the high-performance computing at the University of Oldenburg. Finally, we sincerely thank Luis Miguel Rodriquez-R, Marike Palmer, and the entire SeqCode team for their expert grammatical and taxonomic guidance. This research was supported by funding from the National Science Foundation grants OCE-1538628 (MSR), DEB-2224832 (MSR), and OCE-2149128 (MSR) as well as the Simons Postdoctoral Fellowship in Marine Microbial Ecology (LS-FMME-00989028) (SJT). This is HIMB publication 2025 and SOEST publication 12043.Author informationAuthors and AffiliationsHawai‘i Institute of Marine Biology, School of Ocean and Earth Science and Technology, University of Hawai‘i at Mānoa, Kāne‘ohe, HI, USAKelle C. Freel, Sarah J. Tucker, Evan B. Freel & Michael S. RappéMarine Biology Graduate Program, University of Hawaiʻi at Mānoa, Honolulu, HI, USASarah J. TuckerJosephine Bay Paul Center for Comparative Molecular Biology and Evolution, Marine Biological Laboratory, Woods Hole, MA, USASarah J. Tucker & A. Murat ErenHelmholtz Institute for Functional Marine Biodiversity, Oldenburg, GermanySarah J. Tucker & A. Murat ErenAlfred Wegener Institute, Helmholtz Centre for Polar and Marine Research, Bremerhaven, GermanySarah J. Tucker & A. Murat ErenFort Lauderdale Research and Education Center, Department of Microbiology and Cell Science, Institute of Food and Agricultural Sciences, University of Florida, Davie, FL, USAUlrich StinglDepartment of Microbiology, Oregon State University, Corvallis, OR, USAStephen J. GiovannoniInstitute for Chemistry and Biology of the Marine Environment, University of Oldenburg, Oldenburg, GermanyA. Murat ErenMax Planck Institute for Marine Microbiology, Bremen, GermanyA. Murat ErenAuthorsKelle C. FreelView author publicationsSearch author on:PubMed Google ScholarSarah J. TuckerView author publicationsSearch author on:PubMed Google ScholarEvan B. FreelView author publicationsSearch author on:PubMed Google ScholarUlrich StinglView author publicationsSearch author on:PubMed Google ScholarStephen J. GiovannoniView author publicationsSearch author on:PubMed Google ScholarA. Murat ErenView author publicationsSearch author on:PubMed Google ScholarMichael S. RappéView author publicationsSearch author on:PubMed Google ScholarContributionsK.C.F., S.J.T., A.M.E. and M.S.R. conceived the study, developed methodology and led the investigation and visualization for the study. A.M.E. and M.S.R. supervised the study. KCF wrote the original draft. K.C.F., S.J.T., E.B.F., U.S., S.J.G., A.M.E. and M.S.R. reviewed and edited the manuscript.Corresponding authorCorrespondence to
    Michael S. Rappé.Ethics declarations

    Competing interests
    The authors declare no competing interests.

    Peer review

    Peer review information
    Nature Communications thanks Robin Rohwer, David Walsh, and the other anonymous reviewer 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 InformationDescription of Additional Supplementary FilesSupplementary DataReporting SummaryTransparent Peer Review fileRights 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 articleFreel, K.C., Tucker, S.J., Freel, E.B. et al. New SAR11 isolate genomes and global marine metagenomes resolve ecologically relevant units within the Pelagibacterales.
    Nat Commun (2025). https://doi.org/10.1038/s41467-025-67043-6Download citationReceived: 10 March 2025Accepted: 20 November 2025Published: 14 December 2025DOI: https://doi.org/10.1038/s41467-025-67043-6Share 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