Exploring microbial natural products through NMR-based metabolomics
De-Gao
Wang†,
Jia-Qi
Hu†,
Chao-Yi
Wang†,
Teng
Liu
,
Yue-Zhong
Li
and
Changsheng
Wu
*
State Key Laboratory of Microbial Technology, Institute of Microbial Technology, Shandong University, 266237 Qingdao, P. R. China. E-mail: wuchangsheng@sdu.edu.cn
Received
13th November 2024
First published on 12th June 2025
Abstract
Covering: 2000. 01 to 2025. 03
The soaring demand for novel drugs has led to an increase in the requirement for smart methods to aid in the exploration of microbial natural products (NPs). Cutting-edge metabolomics excels at prompt identification of compounds from complex mixtures and accordingly accelerates the targeted discovery process. Although MS-based metabolomics has become a staple in this field, the utilization of NMR-based metabolomics has severely trailed in comparison. Herein, we summarize the key methodological advancements in 1D and 2D NMR techniques in the past two decades, especially for the invention of computational technologies and/or introduction of artificial intelligence for automated data processing, which significantly strengthen the ability of NMR-based metabolomics to analyze crude microbial extracts. Preliminary fractionation is advocated to deconvolute samples and thus enhance detection sensitivity towards minor components overshadowed by a complex matrix. Particularly, the synergistic application of NMR-based metabolomics and genomics provides an expedient approach to correlate biosynthetic gene clusters with cognate metabolites, greatly improving the efficiency of dereplication and, thus, targeted discovery of novel compounds. A variety of microbial NPs involving distinct chemical skeletons and/or biosynthetic logics are enumerated to prove the genuine prowess of NMR-based metabolomics. Overall, this review aims to encourage the broader adoption of NMR-based metabolomics in the realm of microbial NP research.
De-Gao Wang
De-Gao Wang started his doctoral study and scientific research under the supervision of Prof. Changsheng Wu in 2023. He holds degrees in Biological Medicine (M.S. 2023) from Shandong University. His doctoral research focuses on the discovery of myxobacterial natural products with interesting skeletons and significant biological activity.
Dr Jia-Qi Hu
Dr Jia-Qi Hu obtained her Bachelor's Degree in 2016 and Master's Degree in 2019 from Beijing Forestry University. She studied Microbiology at the Institute of Microbial Technology, Shandong University and received her PhD in 2023. Then, she continued with post-doctoral research. Currently, she is aiming to discover novel bioactive secondary metabolites from myxobacteria through the approach of integrating genome mining and NMR-based metabolomics.
Dr Chao-Yi Wang
Dr Chao-Yi Wang obtained her PhD in Medicinal Chemistry from Ocean University of China in 2018 under the supervision of Prof. Chang-Yun Wang. She started working as a Postdoctoral Fellow at Peking University with Prof. Wen-Han Lin from 2018 to 2021, and now at Shandong University with Prof. Changsheng Wu. Her current research interests focus on the discovery of novel bioactive natural products from myxobacteria.
Prof. Yue-Zhong Li
Prof. Yue-Zhong Li received his PhD at Shandong University in 1993. Since 2000, he was appointed a Professor by Shandong University, and won the Outstanding Youth Fund of the National Foundation of China in 2008. He is mainly engaged in research on the cultivation, ecology, molecular biology, and secondary metabolites of myxobacteria. In recent years, he has utilized the (meta)genome mining approach to study the distribution of cultivable and uncultivable microorganisms in various ecological environments.
Prof. Changsheng Wu
Prof. Changsheng Wu studied Pharmacy for his BSc Degree (2008) and Natural Products Chemistry for his MSc Degree (2011) at Shandong University, China. He obtained his PhD in 2016 at Leiden University, The Netherlands, where he discovered new antibiotics from actinomycetes. In 2017, he conducted postdoctoral research at The Rockefeller University, USA, using the metagenomics tool to find new antibiotics. In 2019, he joined Shandong University as a Full Professor. His current research interests focus on the discovery of novel bioactive natural products from myxobacteria.
1. Introduction
Microbial natural products (NPs) have a proven track record of success in the history of drug discovery by serving as an invaluable source of therapeutic agents and potential drug leads.1 Nowadays, most research groups dedicated to the discovery of microbial NPs maintain in-house culture collections, encompassing diverse strains such as actinomycetes, fungi, cyanobacteria, and myxobacteria, all recognized as prolific sources of bioactive secondary metabolites.2 A central challenge lies in effectively leveraging these invaluable strain libraries. Although traditional activity-based high-throughput screening has historically proven successful, it often results in the re-discovery of known compounds.3 Thus, to enhance research competitiveness and fill the drying drug discovery pipeline, there is a pressing need to develop more innovative and efficient methodological strategies to swiftly uncover new lead structures from microbial resources.
Metabolomics, an analytical tool proficient at chemically scrutinizing biological samples, can quickly discern variations in secondary metabolomes across different bacterial fermentations with the aid of chemometric methods, significantly steering efforts toward potential biomarkers for the discovery of novel compounds and avoiding chemical redundancy in an early stage.4 Mass spectrometry (MS) and nuclear magnetic resonance (NMR) are the two most commonly used technologies for metabolomics, primarily because of their powerful ability to detect individual metabolites in complex mixtures. Although MS-based metabolomics has been widely applied in microbial NP research,5,6 NMR-based metabolomics is much less frequently employed in this field but more in plant or clinical applications.7,8 Considering its popularity in the discovery of microbial NPs, existing reviews overwhelmingly focus on MS-based metabolomics,5,6,9,10 whilst a comprehensive review dedicated specifically to the role of NMR-based metabolomics in this respect remains conspicuously absent to date. Moreover, the integration of genomics and metabolomics for accelerating the discovery of microbial NPs has emerged as a dominant strategy in the post-genomic era. Although numerous reviews detail the synergy between MS-based metabolomics and genomics,11–16 a dedicated overview of the same for NMR-based metabolomics is lacking. Given that no perfect method exists to solve any problem and both NMR and MS have strengths and weaknesses, NMR-based metabolomics holds great promise to provide a useful compliment to MS-based metabolomics for the bioprospecting of taxonomically diverse microbes. In fact, NMR boasts several unique advantages, such as being nondestructive, nonbiased, requiring no sample treatment, no ionization, and free from ion annotation. Thus, there has been a significant increase in recent decades in the use of NMR-based metabolomics, either independently or in conjunction with other disciplines such as MS-based metabolomics, genomics, and artificial intelligence (AI), resulting in the accelerated discovery of novel microbial NPs. Accordingly, a timely consolidation of this progress is crucial to encourage wider adoption of these techniques by the broad range of scientific communities engaged in microbial research.
Although there may be subtle distinctions among the conceptual terms “NMR-based metabolomics,” “NMR-based metabolic profiling,” and “NMR fingerprinting,” we chose to use “NMR-based metabolomics” in the title as a unifying descriptor. This term reflects the shared underlying principle of utilizing NMR technology to analyze the chemical composition of complex mixtures.17–20 Rather than reiterating the fundamental concept of NMR-based metabolomics, which has been extensively covered elsewhere,17,18,21 we structured the content from an application-oriented perspective, with the aim to practically leverage this tool for the discovery of microbial NPs. Therefore, this review is structured into three major sections, each interconnected but with a distinct focus. Section 2 focuses on the general steps and principles of NMR-based metabolomics in microbial NP research. It enumerates the common one-dimensional (1D) and two-dimensional (2D) NMR-based metabolomics approaches, highlighting key advancements and exploring the application of machine learning and/or AI for NMR data processing and analysis. Section 3 emphasizes and addresses the widely acknowledged challenges in current NMR-based metabolomics, including the inherent complexity of metabolic components and limitations in sensitivity. The integration of prefractionation with NMR-based metabolomics alleviates the problematic signal overlapping and the detection sensitivity for minor components can be significantly improved. Recognizing the increasing dominance of genomics-driven microbial NP exploration, Section 4 focuses on the synergistic combination of NMR-based metabolomics and genomics, demonstrating how this integrated approach boosts the efficiency and target-specificity. Each section showcases illustrative examples of novel microbial NPs with diverse chemical scaffolds, unique structural features, and varying biosynthetic pathways, demonstrating the effectiveness and robustness of different NMR methodologies. However, to ensure the adequate illustration of the concepts and/or principles within each section, some compromises were made in example selection, even when certain examples could have served across multiple sections. Section 5 concludes and offers a perspective on the future of NMR-based metabolomics in microbial NP discovery, emphasizing the potential of its integration with multiple disciplines such as MS-based metabolomics, genomics, and AI to further strengthen its capacity. Ultimately, this review aims to encourage the broader adoption of NMR-based metabolomics in the realm of microbial NP research. We hope to inspire researchers to leverage this powerful tool not only for the discovery of novel bioactive compounds, but also the elucidation of complex biosynthetic pathways, and the detailed investigation of microbial chemical ecological interactions.
2. Advancements in NMR-based metabolomics for microbial NP discovery
2.1 General workflow of NMR-based metabolomics for microbial NP discovery
The typical workflow initiates with the extraction of metabolites from microbial cultures, followed by the acquisition of high-resolution NMR spectra. The resulting complex data are then subjected to rigorous processing, involving phasing, baseline correction, referencing, deconvolution, and peak picking. Subsequently, statistical tools are applied to highlight the metabolic differences between samples, identify potential biomarkers associated with desirable traits (e.g., bioactivity), and dereplicate known compounds. These identified biomarkers guide targeted isolation efforts, facilitating the purification of novel compounds for subsequent structural elucidation.8,20,22 Finally, Computer-Assisted Structure Elucidation (CASE) software, often coupled with density functional theory (DFT)-NMR calculations, is increasingly used to automate spectral interpretation and structure elucidation, streamlining the identification of novel microbial NPs (Fig. 1). The great advancements in both hardware and software make NMR-based metabolomics increasingly accessible to a broader range of researchers, further accelerating the pace of microbial NP discovery.
Fig. 1 General workflow of NMR-based metabolomics for the discovery of microbial NPs.
2.1.1 Sampling. Sample preparation is a critical step in NMR-based metabolomics for microbial NP discovery, influencing the quality and interpretability of the resulting data. To realize the high-throughput screening of microbial strain libraries, efficient and reproducible culture conditions are essential. Cultures can be grown using either solid-state or liquid fermentation methods, with careful optimization of factors such as medium composition, incubation temperature, aeration, and cultivation time to maximize metabolite production. When comparing different strains, it is often advantageous to focus on strains with close phylogenetic relationships, which minimizes the inherent differences in secondary metabolite production pathways, facilitating the identification of more subtle, yet potentially significant, variations in metabolite profiles. Variable elicitation strategies, including the addition of specific chemical or physical stressors, may be employed to stimulate the production of secondary metabolites.23 Given the inherent sensitivity limitations of NMR spectroscopy, it is crucial to optimize the fermentation scales to ensure sufficient metabolite concentrations for reliable detection. Biological replicates are essential to account for biological variability and ensure statistical significance in downstream multivariate data analysis. The selection of an appropriate metabolite extraction method is another key consideration. Conventional techniques such as organic solvent extraction (e.g., ethyl acetate, methanol, and acetone) and solid-phase extraction (SPE) can be employed to selectively isolate the metabolites of interest, while minimizing the extraction of interfering composites (e.g. salts, proteins, and other contaminants that can compromise the NMR spectral quality). Different SPE sorbents (e.g., C18, silica, and HP-20 resin, ion exchange resins) can be tailored to selectively extract microbial NPs with specific properties, increasing the efficiency of NMR data analysis.24 Novel extraction methods are gaining traction in the field of microbial NPs. For example, the Müller group pioneered the application of supercritical fluid extraction (SFE) for the extraction of secondary metabolites from myxobacteria.25 Following extraction, the crude extracts are concentrated under vacuum to dryness, and then reconstituted in a deuterated solvent (e.g., typically deuterated methanol) to maximize the sample solubility and provide a field-frequency lock for the NMR spectrometer. Also, parallel extraction of the blank culture medium should also be performed as a control.
2.1.2 NMR data acquisition. High-quality NMR data acquisition is paramount for the successful application of NMR-based metabolomics. This process benefits significantly from advancements in NMR instrumentation. High-field magnets (800 MHz and above) provide increased spectral resolution and sensitivity,26 enabling the detection and characterization of even minor metabolites within complex mixtures. Cryogenic probes, which cool the NMR probe head to extremely low temperatures, further enhance the sensitivity by reducing the thermal noise. Microprobes and capillary probes minimize the sample volume requirements.27 Automated sample handling systems, including robotic liquid handlers and flow injection analysis (FIA)-NMR,28 dramatically increase the throughput, enabling the rapid analysis of hundreds or even thousands of samples. Regarding pulse sequence optimization, 1H NMR spectra are typically acquired using pulse sequences that incorporate water suppression techniques,29 such as WATERGATE,30 to minimize the dynamic range of the water signal and allow the detection of minor metabolites. The number of scans (NS), spectral width (SW), and acquisition time (AQ) are adjusted to achieve the optimal signal-to-noise ratio (SNR) and resolution.31
2D NMR experiments, such as HSQC, HMBC, and COSY, provide valuable structural information for the identification and dereplication of metabolites. However, one of the major drawbacks of 2D NMR experiments is their time-intensive process, typically in the order of 3 to 4 h per sample. Gratifyingly, there has been a surge in new methods for shortening the measurement time of multidimensional NMR spectra, while retaining high spectral resolution along the indirect dimension(s).32 For instance, a number of non-uniform sampling (NUS) approaches bypass the traditional Fourier transformation (FT) processing.33 Besides, the covariance NMR has demonstrated effectiveness by improving the spectral resolution, boosting the sensitivity, shortening the experimental time, and enabling the observation of correlations among less sensitive heteronuclei.34 Absolute minimal sampling (AMS) enables a drastic reduction in the measurement time required for high-throughput metabolomics applications, facilitating the rapid identification and quantification of components within complex biological mixtures.35 NOAH supersequences36 combine multiple conventional NMR pulse sequences, such as HSQC, HMQC, HMBC, COSY, NOESY, and TOCSY, into a single measurement. This approach offers significant time savings and increases the efficiency of NMR measurements, thus dramatically increasing the sample throughput. However, although NUS and AMS can enhance the resolution or processing speed, their advantages in heteronuclear experiments with naturally abundant 13C are somewhat constrained due to inherent sensitivity limitations. To address this problem, Hansen et al. demonstrated that the inherent sensitivity of the HSQC and HSQC-TOCSY experiments can be augmented by incorporating an isotope-selective composite pulse sequence at the onset of the standard sensitivity-boosting protocols. Further incorporation of sensitivity-improved versions of HSQCsi and HSQCsi–TOCSY in NOAH supersequences enabled the compact acquisition of multiple 2D NMR data sets with significant improvements in sensitivity, resolution, and/or time.37 All these (continued) advancements hold great promise to propel the utility of 2D NMR-based metabolomics studies with large cohorts of microbial samples. Collectively, the synergistic combination of advanced instrumentation and optimized acquisition parameters ensures the acquisition of high-quality NMR data.
2.1.3 NMR data processing. Following data acquisition, the efficient and accurate processing of the generated NMR data is paramount.38 Phasing corrects for any distortions in the peak shape caused by instrumental imperfections, ensuring that the peaks are purely absorptive and symmetrical. This is typically achieved through manual or automated adjustment of the zero-order and first-order phase correction parameters.39–41 Baseline correction is applied to remove any unwanted background signals or distortions that may arise from instrumental artifacts or imperfect shimming. This process involves fitting a polynomial or other mathematical function to the baseline region of the spectrum and subtracting it from the entire spectrum, resulting in a flattened baseline and improved peak visibility.42,43 Referencing involves calibrating the chemical shift scale by aligning a known reference signal, such as the residual solvent signal or an added internal standard (e.g., tetramethylsilane), to its established chemical shift value. This ensures accurate and consistent chemical shift assignments across different samples and experiments.44,45 Automated data processing of phasing, baseline correction, and referencing, is indispensable to significantly reduce the need for manual intervention and improve the data reproducibility across experiments.38,46 Specialized software packages, such as TopSpin (Bruker) and NMRPipe, and open-source platforms such as NMRglue47 provide automated routines for these tasks. Effective noise reduction is equally important for extracting meaningful information from complex spectra, and advanced algorithms, such as iPick,48 are increasingly employed to boost the overall spectral quality.
NMR spectral deconvolution is a critical in metabolomics studies, particularly for complex mixtures, given that it resolves overlapping signals to enable accurate the identification and quantification of individual components.49,50 By disentangling composite signals into their constituent peaks, defining their position, intensity, and line shape, deconvolution enhances the resolution of complex spectra, revealing low-abundance metabolites otherwise obscured by signal overlap. This process not only improves the accuracy of metabolite quantification but also simplifies spectral interpretation, ultimately facilitating comprehensive metabolomic profiling and biomarker discovery. Gratifyingly, recent advancements in deep learning algorithms have further enhanced the capabilities of NMR spectral deconvolution,49–51 allowing for the resolution of overlapping signals and the extraction of more detailed information from complex spectra. For example, Schmid et al.52 introduced a novel deep learning-based algorithm for the accurate and automated deconvolution of 1D NMR spectra, including resolving overlapping peaks, handling high dynamic range signals, and maintaining sparsity in the resulting peak lists. Nicolas Schmid and colleagues successfully developed an intelligent deconvolution algorithm based on deep learning, achieving expert-level accuracy in bioinformatics processing.52
Peak picking is another critical process for identifying and selecting effective signals in the NMR spectrum, which directly impacts downstream analysis and metabolite identification.53 The automatic and fast identification of individual peaks in 1D NMR spectra and cross-peaks in 2D or higher dimensional NMR spectra is essential for the high-throughput analysis of complex NMR spectra. Algorithms have been developed to automate this process, and simultaneously minimize the erroneous identification of spurious peaks, ensuring the reliability of subsequent analysis. For instance, Li et al. presented DEEP Picker,53 a DNN-based system for automated peak detection and deconvolution of 2D NMR spectra. This tool, trained on a broad set of synthetic spectra, employs deep convolutional layers to pinpoint even subtle overlapping peaks. It is compatible with platforms such as COLMARq54 to facilitate an integrated metabolomic analysis workflow that simplifies data handling and interpretation. In addition, Li et al. also developed DEEP Picker1D and Voigt Fitter1D for the fully quantitative deconvolution of 1D NMR spectra, greatly streamlining peak picking, fitting, and reconstruction.55
2.1.4 Multivariate data analysis of NMR spectra. Multivariate data analysis (MDV) enables the extraction of meaningful biological insights from complex spectral datasets.56,57 These methods operate by reducing the dimensionality of the data, allowing the visualization of underlying patterns and relationships between samples.58 In NMR metabolomics, this often involves an initial step of “bucketing” or binning the spectra, dividing the chemical shift range into small, equally sized regions and using the integrated signal intensity within each bucket as a variable for subsequent analysis.59,60 Commonly employed multivariate techniques include principal component analysis (PCA), which is an unsupervised method for exploring data structure, and supervised methods such as orthogonal partial least squares discriminant analysis (OPLS-DA), maximizing the separation between predefined groups of samples. A key output of these analyses is the score plot, which displays the relationships among many samples, highlighting outlier groups or clusters with distinct metabolic profiles. Then, the corresponding loading plots reveal the specific NMR signals (biomarkers) responsible for driving these separations, providing valuable clues about the underlying chemical differences. These approaches enable the prioritization of “gifted strains” or the optimization of culture conditions to enhance the production of desired metabolites. The popular software packages for performing these analyses include MetaboAnalyst, SIMCA, and R-based packages such as ropls and ChemoSpec.
2.1.5 NMR interpretation and dereplication. Following multivariate data analysis, a key question arises, how can we identify the compounds responsible for the separation observed in the score plot of MDV? At the crude extract level, dereplication serves as a critical initial step to streamline the discovery workflow by enabling researchers to focus their resources on the identification and characterization of truly novel molecules, thereby minimizing the wasteful re-isolation of previously reported compounds.61–63 This can be accomplished by matching experimental chemical shifts or peaks against NMR metabolomics databases64 (e.g. BMRB,65 HMDB,66 and DEREP-NP63). However, most available NMR metabolomics databases are manually curated for primary metabolites and not suitable for structurally diverse secondary metabolites.67,68 Therefore, the practical dereplication of known specialized metabolites in microbial crude extracts often necessitates a combination of 1D and 2D NMR experiments, sometimes with assistance from mass spectrometry (MS). Then, the generated complete or partial structures are used as “probes” to query against natural product databases, for instance, by performing substructure or similarity searches utilizing the predicted chemical structures. Numerous databases have been developed, ranging from those dedicated to specific microbial classes to more inclusive repositories. Examples include MyxoDB for myxobacterial NPs,69 StreptomeDB for compounds derived from Streptomyces,70 CyanoMetDB for cyanobacterial secondary metabolites,71 and comprehensive databases such as The Natural Product Atlas72 and the Natural Product Dictionary, which catalog NPs produced by a wide variety of microorganisms. These dedicated or comprehensive databases serve as invaluable resources for rapidly identifying and avoiding known compounds from further consideration, allowing researchers to focus their efforts on the discovery of truly novel microbial NPs. An excellent review summarizing natural products databases was published elsewhere.73 Notably, although the generated structural fragments at the crude extract level may not always be sufficiently unique for direct database matching and dereplication, experienced natural product chemists can often integrate NMR data with their biosynthetic knowledge to infer the general compound class (e.g., terpene, macrolide, aromatic polyketide, and polypeptide) represented by the structural fragment. This crucial information provides invaluable guidance for subsequent NMR-guided separation, particularly in selecting appropriate chromatographic methods to facilitate targeted isolation. For instance, the co-existence of the structural fragments of oxygenated methines (δH 3.5–4.8), aliphatic methyl groups (δH 0.5–1.5), and olefins (δH 5.0–6.0) in an NMR profile suggests the presence of the macrolide superfamily of microbial NPs.
2.1.6 NMR-guided separation for de novo structural identification. However, in most cases, the novel compounds corresponding to the biomarkers identified by MDV are present in relatively low concentrations within the crude extract, making their direct identification impossible at the crude extract level. Alternatively, it is sometimes challenging to dereplicate smoothly because the 2D NMR-resolved structural features may not be unique, and thus many different hits are retrieved from databases. Therefore, NMR-guided isolation of biomarkers is inevitable for de novo structure elucidation. This process typically begins with scaling up the “outlier groups” identified by MDV. It is highly recommended to maintain exactly the same culture conditions (media, temperature, aeration, and elicitation) and extraction method used during the initial high-throughput screening to ensure metabolic reproducibility. The, the crude extracts are subjected to a series of chromatographic separations, including normal-phase chromatography, size-exclusion chromatography, and reversed-phase HPLC. During each separation step, the fractions are meticulously collected and analyzed by NMR spectroscopy, focusing on the characteristic NMR signals of the target metabolites identified through prior MDV. Fractions exhibiting the desired signals are selectively pooled, while those lacking the target signals are discarded, enabling the efficient removal of unwanted compounds and the enrichment of metabolites of interest. Once a pure compound is obtained, comprehensive structural characterization is performed. This typically involves a suite of 1D and 2D NMR experiments (1H NMR, 13C NMR, COSY, HSQC, and HMBC), complemented by high-resolution MS for molecular weight determination. Nuclear Overhauser effect (NOE) or rotating-frame Overhauser effect (ROE) experiments are employed to define spatial relationships between protons, and J-coupling analysis is performed to determine dihedral angles.74,75 Together, these data sets provide the relative stereochemistry. For absolute configuration determination, the experimental electronic circular dichroism (ECD) spectra are compared with calculated spectra,76 or chiral derivatizing agents (e.g., Mosher's reagent77 and Marfey's reagent78) are used to generate diastereomers, allowing comparative NMR analysis. In favorable cases, single-crystal X-ray diffraction may be employed to confirm the planar skeletons and/or to determine the stereochemistry. Recent advancements in non-spectroscopic methods, including atomic force microscopy (AFM),79 “crystalline sponge” X-ray analysis,80,81 Ag3Pz3 crystalline mate,82 and micro-electron diffraction (micro-ED),83 are providing powerful assistance for structure elucidation, particularly for compounds with limited material or those that form poorly diffracting crystals. These techniques offer complementary information to traditional methods, expanding the scope of structural characterization in chemical research. Finally, the isolated compound is subjected to bioactivity testing, validating the initial MDV results and confirming its potential as a novel microbial NP.
2.1.7 Computer-assisted structure elucidation (CASE). The CASE software streamlines de novo structure elucidation, which has been intensively reviewed elsewhere.84,85 Programs such as ACD/Structure Elucidator and Bruker CMC-se86 leverage algorithms to generate structures consistent with the input data (molecular formula, chemical shifts, and 2D NMR correlations). Algorithms analyze COSY, HSQC, and HMBC data to establish atom connectivity within the molecule, constructing a molecular connectivity diagram as a foundation for structure generation. Although the traditional CASE methods focus primarily on planar structure elucidation, the integration of parameters derived from residual dipolar couplings (RDCs),87 residual chemical shift anisotropies (RCSAs),88 and coupling constants (nJHH, nJCH, and nJNH)89 now enables the generation or assignment of stereochemistry for novel structures. Furthermore, DFT-NMR calculations are being increasingly used, both for determining 3D structures and estimating the 13C and 1H chemical shifts of alternative structures for comparison with experimentally derived chemical shifts. Nevertheless, DFT-based NMR chemical shift calculations tend to be computationally demanding, and rapid data-driven approaches frequently fall short in terms of reliability. Thus, to overcome this dilemma, emerging tools such as DU8ML,90,91 ML-J-DP4,92 GFN2NMR,93 and SVM-M,93 combine quantum mechanical precision with machine learning (ML)-driven pattern recognition, enabling the rapid resolution of stereochemical complexities.92 For instance, ML-J-DP4 integrates 3JHH couplings with ML optimization, accelerating structural elucidation 100-fold and resolving the stereochemistry of belizentrin methyl ester (32168 candidates) within 90 min.92 The Wang group developed GFN2NMR,93 a deep graph convolutional network (GCN), for calculating 13C NMR chemical shifts with near-DFT accuracy, while maintaining low computational cost. The model demonstrated high performance in both general 13C NMR prediction and challenging structure assignment tasks, including elucidating the J/K ring junction stereochemistry of maitotoxin, the largest known non-polymeric natural product. The continued development and refinement of CASE tools are poised to further promote NMR-based metabolomics, making it more accessible and widely applicable in the field of microbial NP discovery.
2.2 One-dimensional NMR-based metabolomics applied for microbial NP discovery
2.2.1
1H NMR-based metabolomics. 1H atoms are present in nearly all organic compounds and produce the highest NMR signal intensity, and thus 1H NMR is oftentimes chosen as the preferred method due to its relatively short acquisition times (typically just a few minutes for one sample) compared to 13C or 2D NMR experiments. The signals from proton resonances are directly proportional to the molar concentration of the substances, facilitating the straightforward comparative analysis of different samples. Proton chemical shifts are acutely responsive to variations in pH, temperature, salinity, and the chemical makeup of the metabolomics sample. Furthermore, advanced NMR techniques, including PSYCHE (pure shift yielded by chirp excitation) and Zangger–Sterk experiments,94 enhance the spectral resolution, enabling the discrimination and identification of closely spaced peaks and facilitating more accurate and comprehensive compound annotation. However, owing to its limited spectral range (typically 0–12 ppm), the 1H NMR spectroscopy suffers from severe signal overlap, which confuses data interpretation especially in the case of complex mixtures. Therefore, 1H NMR-based metabolic profiling is commonly used together with multivariate data analysis (e.g. PCA, PLS-DA, and OPLS) to deconvolute complex mixtures for biomarkers identification.7 The benefits of this approach are two-pronged; firstly, it enables the rapid and straightforward identification of known compounds within mixtures, eliminating the need for additional isolation steps; secondly, it facilitates the detection of signals that deviate from those of known compounds, thereby aiding in the discovery of novel molecular entities.
In our 1H NMR-based metabolomics workflow for the bioprospecting of actinomycetes, antibiotic production by microbial NP producers was first perturbed by different eliciting strategies, such as microbial cocultivation,95 streptomycin-resistance mutation,96 overexpression of pathway-specific regulator,97 and varying the harvest time98 or culture media.99 Subsequently, chemometric methods were used to deconvolute the complex 1H NMR metabolic profile and compare the spectra of different experimental groups, accordingly illustrating the unique biomarkers that correlate with bioactivities. The biomarker-oriented structural elucidation was conducted based on 2D NMR experiments including 1H–1H COSY, 1H–13C HSQC, HSQC–TOCSY, 1H–13C HMBC, and J-resolved NMR to identify associated structural fragments. 1H NMR-guided separation by tracking the biomarker signals resulted in the discovery of tri-methoxylated isocoumarins (1),98 7-prenylisatin (2),96 aromatic polyketide GTRI-02 (3),95,100 and naphthoquinone juglomycin C amide (4).99 In addition to actinomycetes, we also leveraged 1H NMR-based metabolomics for investigating myxobacteria. We constructed the first publicly accessible database, MyxoDB (https://www.myxonpdb.sdu.edu.cn), which is solely dedicated to myxobacterial NPs.69 It greatly simplified the dereplication of commonly known myxobacterial compounds (e.g. phenalamide, myxothiazol, and corallopyronin) by analyzing the 1H NMR spectra of 147 strains from our in-house myxobacterial collection. Further PCA analysis could distinguish the outlier strain, Archangium gephyra SDU49, and the signals responsible for its distinct separation were pinpointed. NMR-guided separation by tracking the biomarker signals enabled us to swiftly identify a class of linear di-lipopeptides named archangiumins (5–8).
Beyond our own investigations, Betancur et al. also demonstrated the utility of 1H NMR-based metabolomics in the discovery of novel microbial NPs. In their study, Streptomyces sp. PNM-9 was cultivated in four distinct liquid media for 15 days to examine the influence of the growth medium composition on the production of bioactive metabolites. Samples were collected at specified time points (days 1, 2, 3, 4, 5, 10, and 15) and subjected to 1H NMR-based fingerprinting to profile the resulting metabolic landscapes. Subsequent OPLS-DA analysis was employed to correlate the obtained 1H NMR spectral data with observed bioactivity. Targeted isolation guided by biomarker identification from the OPLS score plot led to the characterization of 2-methyl-N-(2-phenylethyl)-butanamide (9) and 3-methyl-N-(2-phenylethyl)-butanamide (10).
2.2.2
13C NMR-based metabolomics. In comparison to 1H NMR spectroscopy, the application of 13C NMR-based metabolomics for microbial NP research has largely lagged behind. This is primarily due to the low natural abundance of the 13C isotope (1.1%) and its low gyromagnetic ratio (only 25% that of 1H), which significantly diminishes its detection sensitivity. However, the advancements in high-magnetic field NMR spectrometers and cryogenic probes now allow the acquisition of 13C NMR spectra with high resolution and sensitivity in a relatively short time frame. Wu et al. developed a deep neural network named DN-Unet,101 a data postprocessing technique to increase the signal-to-noise ratio of liquid-state 13C and multi-dimensional NMR spectra in a cost-effective way. This technology has been proven to be effective for denoising and adeptly preserving weak peaks that are typically obscured by noise, while effectively quelling spurious peaks in the 13C NMR spectra. In fact, 13C NMR offers significant benefits for analyzing complex mixtures because it provides a direct assessment of the metabolite backbone structures. Each carbon in a molecule typically corresponds to a unique resonance in a 1H-decoupled 13C NMR spectrum. Moreover, the spectral width of 13C NMR is substantially broader than that of 1H NMR (220 ppm for 13C versus 12 ppm for 1H), which greatly minimizes the occurrence of signal overlaps. On account of these advantages, the Renault group devised a dereplication strategy that integrates 13C NMR-based metabolic profiling with multigram-scale fractionation and hierarchical cluster analysis (HCA) for the rapid identification of major metabolites in natural extracts.102 This strategy involves fractionating a mixture, followed by 13C NMR analysis of the resulting fractions.103 Automated gathering and alignment of the 13C signals across various spectra, followed by HCA allows the identification of correlations between signals from the same molecular structure within the diverse mixtures, manifesting as chemical shift clusters. Matching experimental 13C NMR data in each cluster to a 13C NMR chemical shift database reveals specific molecular structures. The prefractionation step can be omitted by introducing a computational database search algorithm.104 In a separate report, 13C NMR data of superior quality were obtained utilizing a specifically designed 13C-optimized probe and used to establish 13C–13C statistical correlations. This significantly enhanced the accuracy of metabolite identification under natural abundance conditions.105 The Richomme group developed MixONat,106,107 a tool that compares the 13C NMR spectroscopic data from a raw extract against either theoretical predictions or experimental data stored within databases. A distinctive aspect of MixONat lies in its implementation of carbon multiple filtering, a critical phase that markedly improves the accuracy of matching. Furthermore, the integration of MixONat based on 13C-NMR with 2D NMR profiling has demonstrated efficacy in swiftly dereplicating and pinpointing new secondary metabolites within exceedingly complex mixtures.106–108 Altogether, advances in instrumentation and data processing are positioning 13C NMR-based metabolomics as a promising, albeit underexplored, avenue for microbial NP discovery due to its ability to provide unique and valuable structural insights.
2.3 Two-dimensional NMR-based metabolomics in microbial NP discovery
The restricted spectral dispersion and complex signal patterns in the 1D NMR spectra of crude extracts result in substantial signal overlap, hindering their utility in dereplication. Conversely, 2D NMR techniques, providing expanded spectral dispersion and additional orthogonal dimensions, minimize signal overlap in complex mixtures and yield more comprehensive structural insights. In theory, all commonly used 2D NMR experiments, such as 1H–1H COSY, 1H–1H TOCSY, 1H–13C HSQC, and 1H–13C HMBC, can be employed to extract the necessary structural information for identifying chemical entities in mixtures.
2.3.1 HSQC. HSQC is notably sensitive, providing connectivity data for directly bonded 1H and 13C (or 15N) nuclei. The HSQC method is extensively utilized for metabolomics study owing to its ability to generate high-resolution spectral data by correlating carbon and proton signals within a molecule. Several computational tools have been developed to facilitate HSQC-based metabolomics studies. Duggan et al. developed an “atomic sort” method (AFM) for the HSQC spectral analysis of mixtures.109 They built an HSQC database using public spectra from the Human Metabolome Database66 and BioMagResBank database,110 identifying 10308 1H–13C HSQC correlations from 1207 spectra as typical molecular features. Euclidean distances were calculated between sample peaks and the database, with outliers indicating potential unknown compounds for chromatographic isolation. Consequently, this method was applied to a mere 52 μg of ethyl acetate extract from marine-derived Streptomyces sp. CNB-982. A clear spectroscopic beacon signal (δH −0.90 and δC 32.3, distance 13.74%) was highlighted, guiding the targeted chromatographic separation. This enabled the identification of the antibacterial cyclic heptapeptide cyclomarin A (11).111 Although cyclomarin A was originally discovered in 1999, it still remained a compelling example of the capacity of AFM to reveal rare structural features at the crude extract level to speed up the targeted discovery process. Indeed, the authors subsequently applied the AFM methodology to 32 μg of crude extract from the marine sponge Plakortis halichondrioides, which led to the discovery of a new compound, gracilioether L (12). SMART (Small Molecule Accurate Recognition Technology), initially presented in 2017, employs a deep convolutional neural network (CNN) to automate the identification and annotation of compounds from HSQC spectra.112 SMART outputs a curated list of probable compound structures, leveraging a CNN model trained on a vast array of HSQC spectra from characterized natural molecules, 2054, in the original version and an expanded 53076 in SMART 2.0.113 In 2020, SMART 2.0 was first applied to mixture analysis, successfully predicting the presence of a novel macrolide termed symplocolide A (13) in an active fraction of the crude extract originating from the filamentous marine cyanobacterium Symploca sp., as well as fast dereplication of known compounds such swinholide A, samholides A. The researchers further compared the capabilities of SMART 2.0 and AFM in identifying rare structural motifs. When SMART 2.0 was applied to the HSQC spectra of pure cyclomarin A and Streptomyces sp. CNB-982 crude extract, all the top hits (2 for cyclomarin A and top 20 for the extract) were closely related ilamycins that harbor the rare N-(1,1-dimethyl-2,3-epoxypropyl)-tryptophan moiety, as also found in cyclomarin A, highlighting its ability to detect structural similarity based on chemical shift patterns, even without directly measuring heteroatoms or quaternary carbons. In effect, the rapid search, taking less than 30 min (structure prediction at 8 s) from spectra to data, demonstrates the efficiency of SMART 2.0. In contrast, three out of the four “novel features” identified by AFM method in cyclomarin A were found to be false positives regarding structural novelty. DeepSAT114 is another recently developed cheminformatic tool that leverages a CNN to directly extract chemical features from HSQC spectra, enabling molecule identification even without authentic NMR spectra. It is trained on a large and diverse dataset of both literature and computationally derived HSQC spectra, and it predicts chemical fingerprints, molecular weights, and structure classes. Compared to SMART 2.0, DeepSAT demonstrated a significant improvement in performance. Specifically, DeepSAT achieved over a two-fold higher correct identification rate than SMART 2.0 (41.0% vs. 18.7% in the top 1 output). Furthermore, unlike SMART 2.0, which does not support multiplicity-edited HSQC data, DeepSAT can be trained with both data to increase the prediction accuracy, with the latter reaching 45.2% and 64.9% in the top-1 and top-5 identifications, respectively. Also, SMART 2.0 relies on a limited in-house database (Moliverse, ∼130000 compounds), whereas, in principle, DeepSAT can be extended to search vast public databases through its structure prediction and fingerprint generation capabilities.
2.3.2 HMBC. HMBC offers the crucial ability to identify correlations between 1H and 13C atoms across distances of two, three, or even four chemical bonds, which is a greater advantage for detecting the positions of quaternary carbons. Bakiri et al. devised an HMBC-based dereplication method that employs a networking strategy to decipher complex NMR spectra of metabolite mixtures. The approach involves generating a theoretical HMBC correlation database from predicted chemical shifts of natural products. Then, a network of experimental HMBC correlations is constructed to identify molecular patterns or structures. The identification of metabolites is facilitated by aligning theoretical HMBC correlations with that from network-analyzed clusters.115 MatchNat, which is predicated on the pattern recognition of specific signals from HMBC spectroscopy, facilitates the swift identification of natural product subtypes.116 Nonetheless, the broad application of HMBC in metabolomics is hindered by its low sensitivity, spectral crowding and overlap, deterioration of spectral quality, and wide concentration range affecting the dynamic range. The powers group enhanced metabolite identification in complex biological samples by integrating HMBC, HSQC, and HSQC-TOCSY data. They adapted a 13C-decoupled HMBC sequence for 13C-enriched samples and compiled a reference database of HMBC spectra for 94 common metabolites, thus bolstering the assignment reliability. Automated assignment was further enhanced by combining a13C–13C covariance matrix from the 13C-decoupled HMBC experiment with the HSQC–TOCSY spectrum.117
2.3.3 HSQC–TOCSY. HSQC–TOCSY NMR experiments combine the benefits of TOCSY and HSQC, providing detailed carbon and proton correlations that link protons to their directly attached carbons and within the same spin system. This makes HSQC-TOCSY highly effective for analyzing complex molecules, particularly those with overlapping 1H and 13C signals.118 MADByTE119 is a novel open-source, NMR-based metabolomics platform designed for the untargeted analysis of complex natural product mixtures. By integrating 1H–13C connectivity data from HSQC spectra with 1H–1H scalar coupling information from TOCSY spectra, MADByTE identifies spin system features within individual mixtures and matches these features across samples to construct chemical similarity networks of samples and shared spin systems (Fig. 2). When samples are subjected to NMR analysis and biological activity assays, samples with promising biological responses can be highlighted, and their characteristic spin systems observed. This approach enables the dereplication of known compounds, the prioritization of bioactive constituents, and the grouping of complex mixtures based on shared compound families, all directly from NMR data without requiring pre-existing spectral libraries. Carroll et al. developed an HSQC–TOCSY fingerprinting approach to prioritize 119 ascidian-associated actinomycetes for the discovery of polyketides and peptides.120,121 When applied to crude extracts, HSQC–TOCSY was found to be significantly more sensitive than HMBC experiments, and effective to determine common structural fragments such as oxygenated methines and indole group, which are commonly associated with polyketides and peptides. This enabled the dereplication of the known macrolide glycoside elaiophylin (14) and indole-containing diketopiperazine naseseazine A (15) in the crude extracts without any fractionation,120 and subsequently resulted in the discovery of a new antiplasmodial polyketide, herbimycin G (16).121
Fig. 2 Schematic of the MADBbyTE workflow. HSQC and TOCSY experiments are performed to acquire 2D NMR data. Subsequently, the spin system is constructed by integrating HSQC-derived 1H–13C correlations with TOCSY-derived 1H–1H coupling networks. After extracting all resolvable spin systems from the samples, similarity calculations and correlation analysis on the spin systems are performed for each extract. By constructing a spin system similarity network, topologically connecting nodes representing spin systems with similar chemical shift patterns, a visualized network model with dual analytical values is generated. The clustering of repeated nodes in the network directly reveals structural homogeneity among extracts, providing structural evidence for the rapid elimination of redundant information. Isolated nodes or weakly connected sub-networks located at the topological periphery of the network often correspond to unique spin system arrangements in the extracts, indicating the presence of novel compound frameworks or rare substituent groups.
2.3.4 Others. In cases where metabolite concentrations are too low for detection in HSQC and/or HMBC spectra, homonuclear 2D 1H–1H TOCSY experiments are a suitable alternative. Distinct from the 1H–1H COSY spectrum primarily reflecting 3J coupling, TOCSY correlates all hydrogen nuclei within the same spin system, thereby providing comprehensive spin–spin connectivity data. DemixC122 was developed by the Brüschweiler group for the interrogation of TOCSY based on covariance processing and deconvolution techniques to cluster rows of spectra. Then, the obtained spectra are matched against customized TOCSY databases,67 thereby aiding in the identification and structural elucidation of unknown mixture components. They also developed a constant-time 13C–13C TOCSY 2D NMR protocol that allowed the de novo determination of carbon backbone topologies of individual metabolites in complex metabolomes,123 enabling the characterization of 112 topologies of unique metabolites in a uniformly 13C-enriched E. coli lysate without any fractionation.124
2D J-resolved (J-Res) spectroscopy separates coupling and chemical shift data across two axes, simplifying signal identification and multiplet analysis. It complements 1D NMR in metabolomics by resolving overlapping peaks more efficiently than most other 2D NMR experiments. Zeng et al. developed the CASCADE experiment for facile multiplet analysis and selective coupling measurement,125 surpassing traditional 2D J-resolved experiments in complex samples with extensive homonuclear couplings. SpinCouple is an online platform utilizing 1H–1H J-resolved spectroscopy and a database of nearly 600 metabolites to identify and annotate metabolites with minimal HSQC peak counts.126J-RESRGAN is a GAN specifically trained to enhance the peak resolution in metabolomic J-Res spectra.127 In the early 2000s, Gerwick et al. introduced DECODES and HETDECODES NMR methods to differentiate species by their diffusion coefficient, enabling the cursory structure elucidation and/or dereplication of the known peptidic compound symplostatin (17) in a complex mixture.128 Schroeder et al. employed double quantum filtered correlation spectroscopy (DQF-COSY) to examine a library of fungal extracts derived from an insect-pathogenic Tolypocladium cylindrosporum strain, which quickly revealed two new terpenoid indole alkaloids, 18 and 19.129 Forseth et al. applied dqfCOSY spectra to compare human pathogen Aspergillus fumigatus with its ΔgliZ mutant, a knock-out strain devoid of the gene encoding the transcriptional regulator of the gliotoxin biosynthetic gene cluster. The metabolomes of ΔgliZ mutant and wild-type strains were individually separated into three fractions to reduce the complexity and polarity range. Differential analysis of the obtained metabolome pools using dqfCOSY profiling allowed a detailed inventory of nine gli-derived metabolites (20–28) featuring unexpected structural motifs.130 In this regard, combined genetic manipulation and differential 2D NMR analysis offer a powerful, unbiased approach for characterizing small-molecule mixtures.
2.4 Specialized heteronuclear NMR techniques
Although microbial NPs are predominantly composed of the chemical elements H, C, O, N, and S, the incorporation of other elements, such as B,131 P,132,133 As,134–137 and Se,138–141 into organic molecules via covalent or coordination bonds can confer unique physiological or pharmacological activities. This has fueled great interest in the targeted discovery of microbial NPs containing these less common elements. In addition to the widely used 1H NMR and 3C NMR, other heteronuclear atoms-based NMR techniques, such as 15N NMR, 19F NMR, 11B NMR, and 31P NMR, can be valuable tools for the straightforward identification of microbial NPs containing specialized functionalities in complex crude extracts, effectively expanding the scope of NMR-based metabolomics.
2.4.1
15N NMR-based metabolomics. Nitrogen is a ubiquitous element in natural products, found in diverse compound classes such as alkaloids, peptides, and various heterocyclic scaffolds. 15N NMR offers unique advantages in directly probing the nitrogen atoms within these complex molecules, providing complementary structural information that augments data from traditional 1H and 13C experiments for characterizing nitrogen-containing microbial NPs. The 15N nucleus exhibits a broad chemical shift range, typically −400 to +200 ppm,142,143 enabling the differentiation of various N-containing functional groups. Chemical shifts are highly sensitive to the electronic environment around the nitrogen atom, permitting the identification of amides, amines, imines, and a variety of N-containing heterocycles. Furthermore, heteronuclear coupling constants, particularly between 15N and directly bonded protons (1JNH) or carbons (nJNC), provide valuable insights into the connectivity and skeleton.144,145 For example, 1H–15N HSQC and HMBC experiments can be used to establish direct and long–range correlations between nitrogen atoms and neighboring protons or carbons, enabling the mapping of complex molecular frameworks. Due to the low natural abundance (0.37%) and negative gyromagnetic ratio of 15N, isotopic enrichment is often necessary to enhance the sensitivity, making it well-suited for targeted metabolomics studies involving specifically labeled precursors or isotopically enriched growth media. The detailed examples of this can be found in Section 4.3.
2.4.2
11B NMR-based metabolomics. Boron is a fascinating element with remarkable chemical and biological properties due to its unique electronic structure. There are a growing number of examples showing that the incorporation of boron into small molecules results in important biological and biomedical signatures.146 Boronic NPs are difficult to identify by mass spectrometry due to their poor ionization and lack of an easy to identify isotope pattern in targeted metabolomics. In addition, boron-containing molecules have no unique identifiers in 1H NMR or other spectroscopic signatures. The 11B NMR chemical shifts, typically ranging from −100 to +100 ppm,147 can provide structural information on boronic acids and related compounds, presenting insight into their coordination environment and bonding characteristics. The MacMillan group significantly propelled this field by employing 11B NMR spectroscopy for both the screening and structural characterization of boron-containing NPs (Fig. 3).148,149 They developed a composite pulse sequence (zgbs) for effective noise suppression to address the challenge of detecting low concentrations of boronated metabolites amidst interfering background signals. The resulting sensitivity-enhanced 11Bzgbsig experiment effectively identified boron-containing metabolites in 13/50 microbial crude extracts, suggesting a higher prevalence of these compounds than previously recognized due to limitations in existing detection methods.148 Accordingly, they prioritized a crude extract from Streptomyces radiopugnans strain SNE- 036, which presented a clear 11B signal at δB 7.0 ppm, a chemical shift suggesting a borate species. Using the 11B NMR-guided separation, they successfully obtained ∼1 mg of pure compound 29 (diadenosine borate). Furthermore, they optimized the coupling parameters for the 1H–11B HMBC experiments to unambiguously elucidate 29 as a novel boron-containing nucleoside dimer.149 Thus, this team stated that their approaches show great promise to characterize previously inaccessible and largely underestimated boron-containing microbial NPs.
Fig. 3 11B NMR-guided discovery of boron-containing NPs. A new pulse sequence (zgbs) can effectively suppress noise and improve the resolution in the 11Bzgbsig NMR experiment. This technique was used to analyze microbial crude extracts for the detection of boron-containing NPs. For example, the clear 11B signal in S. radiopugnans SNE-036 extract-guided 11B NMR-directed separation yielded diadenosine borate (29).
2.4.3
31P NMR-based metabolomics. Phosphorus-containing natural products are a fascinating class of compounds characterized by the presence of a carbon-phosphorus (C–P) bond or other phosphorus-containing functionalities in their molecular structure. These compounds exhibit a diverse range of biological activities, including antibiotic, antiviral, herbicidal, and enzyme inhibitory properties150. 31P NMR is particularly informative for organophosphates, with chemical shifts typically ranging from 0 to −30 ppm for phosphates and up to 250 ppm for phosphonium compounds.151,152 The relatively high natural abundance of 31P (100%) and its good NMR sensitivity make it amenable to direct detection in complex biological samples, revealing key metabolic insights into phosphate metabolism. For instance, Ju and coworkers153,154 employed 31P NMR-based metabolic profiling to facilitate the detection of the chemispecific carbon–phosphorus bond in the complex secondary metabolomes of bacterial strains, which streamlined the targeted discovered novel phosphonic acid NPs such as argolaphos (30 and 31), valinophos (32), (hydroxy)phosphonocystoximates (33 and 34), H-phosphinates (35 and 36), (hydroxy)nitrilaphos (37 and 38), and phosphonoalamides (39–42). Alternatively, 31P NMR-based metabolic profiling tool was leveraged and synthetic biology engineering (heterologous expression, promoter refactoring, gene knockout, in vitro enzymatic reaction, etc.) to investigate the biosynthetic pathway of valinophos (32),155 phosphonothrixin (43),156 and glyphosate (44),157 respectively. 31P NMR spectroscopy plays a crucial role in monitoring and/or quantifying the phosphorus-containing biosynthetic intermediates in complex mixtures.
3. NMR-based metabolomics meets prefractionation
3.1 General principle for concatenated prefractionation and NMR-based metabolomics
Microbial crude extracts are intricate blends, where each constituent has the potential to interact with the biological target under examination. However, only a tiny fraction of components is discernible and measurable by NMR fingerprinting because some may be present at concentrations too minuscule. Additionally, the signal of the bioassay can be obscured by interference from irrelevant or disruptive compounds or by the combined influence of multiple compounds, whether they act additively or synergistically.158 The substantial overlap of signals represents another fundamental obstacle that restricts the broader utilization of NMR-based metabolomics. The chromatographic prefractionation can effectively tackle these problems and strengthen the utility of NMR-based metabolomics. Generally, the strategic workflow (Fig. 4) involves the construction of microbial fraction libraries,159–161 followed by high-throughput screening using NMR-based metabolomic profiling. The development of microbial fraction libraries necessitates meticulous attention to various factors to maximize the structural diversity, bioactivity representation, and overall screening efficiency by NMR. The process commences with the cultivation of phylogenetically diverse microbial strains, followed by the recovery of metabolites to afford crude extracts. A key consideration in terms of the fermentation scale of each strain and/or conditions is balancing the need for sufficient sample material to acquire high-quality 1D and/or 2D NMR spectra and comprehensive bioactivity assay for each fraction, while maintaining a reasonable workload for high-throughput NMR-based screening. In principle, any chromatographic method with sufficient resolving power can be employed for fractionation, including normal-phase silica, size exclusion gels such as Sephadex LH-20, ion exchange resin, and reversed-phase C18 chromatography.162 Given the demands for high-throughput characterization, it is crucial to rationally select chromatographic methods that emphasize ease of operation, high resolving power, and minimal irreversible adsorption. Typically, reversed-phase C18 chromatography utilizing a methanol/water system with a step-gradient elution in 20% increments is empolyed.69 The initial fractions (0% methanol eluent) rich in highly polar medium components are deliberately excluded from further consideration. Therefore, each crude extract is partitioned into five fractions according to polarity, which not only ensures experimental reproducibility but also facilitates the direct NMR comparison of fractions eluted under identical conditions across different strains. The fractionation process can be efficiently performed using automated flash chromatography systems. With the simplified fractions in hand, the NMR-based metabolomic profiling workflow outlined in Fig. 1 can be effectively employed to excavate bioactive microbial NPs. The reduced complexity of the individual fractions significantly enhances the speed and accuracy of both metabolite identification and dereplication. Importantly, given that NMR is an insensitive but non-destructive technique, it is advisable to collect NMR data of fractions prior to bioactivity assays (e.g. antimicrobial, cytotoxicity). Multivariate statistical analyses (PCA, PLS-DA, OPLS-DA, etc.) can be employed to identify fractions with unique or desirable spectral features, and more importantly correlate and these features with bioactivity data. The biomarkers indicative of high activity are used as probes for the NMR-guided separation of bioactive compounds harbored in the fractions of interest. Therefore, prefractionation can improve the targeted discovery efficiency of NMR-based metabolomics.
Fig. 4 Schematic representation of the concatenated prefractionation and NMR-based metabolomics for bioactive microbial NP discovery. A microbial crude extract library is initially subjected to reverse-phase C18 column fractionation using stepwise methanol gradients (20%, 40%, 60%, 80%, and 100% MeOH) in water. Compounds of disparate polarity are separated into different fractions. Although the NMR profiling of crude extracts exhibits high complexity with severe signal overlap, the chromatographic fractionation efficiently reduces the complexity, enhances the sensitivity, and resolves the minor metabolites in the NMR spectra. After NMR measurement, each fraction is subjected to bioactivity assays. Subsequently, multivariate data analysis is employed to discriminate samples and correlate NMR spectral data with bioactivity to identify biomarker signals. Finally, NMR-guided isolation targets these biomarkers, enabling the identification and validation of novel bioactive microbial NPs.
The Quinn group assembled a collection of lead-like enhanced (LLE) fractions,163–166 which were selected due to their advantageous physicochemical characteristics and molecular weight.167 They illustrated the effectiveness of 1H NMR fingerprinting on fractions to discover new substances from a collection of 20 sponges belonging to the Poecilosclerida order. The identification of a distinctive 1H NMR spectral pattern in 5 out of the 220 fractions facilitated the isolation of the novel compound iotrochotazine A (45), which was later found to exhibit phenotypic effects on cells derived from individuals with Parkinson's disease.163 They further applied the same strategy to explore the actinomycetes associated with termite gut, resulting in the discovery of six new compounds, including actinoglycosidines A and B (46 and 47), actinopolymorphol D (48), and niveamycins A, B, and C (49–51).166 Recently, our research group similarly integrated prefractionation with 1H NMR fingerprinting for the discovery of new compounds from myxobacteria. The crude extract of Corallococcus sp. SDU284 was subjected to chromatography on a reverse-phase C18 column, and each resulting fraction indeed gave a simplified 1H NMR fingerprint in contrast to the crude extract. A group of olefinic signals that was not visible in the crude extract became evident in the 1H NMR spectrum of one faction, directing the targeted isolation of corasterol (52), a rare rearranged sterol with noncanonical angular ring fusion pattern.69 In the ELINA workflow, bioactive crude extracts are fractionated to vary chemical compositions and bioactivities. Then, the 1H NMR data of each fraction are correlated with bioactivity using statistical heterocovariance analysis (HetcA) to distinguish biomarkers from active components. When this method was applied to Fomitopsis pinicola, it led to the identification of lanostane triterpenes that inhibit steroid sulfatase.168 In addition to 1H NMR fingerprinting, prefractionation can also work together with 13C NMR or 2D NMR profiling for the targeted discovery of new compounds, as mentioned in Sections 2.1.2 and 2.2.5.102–104,130 Therefore, the synergy between fraction libraries and NMR-based metabolomics improves the targeted efficiency discovery of bioactive microbial NPs.
3.2 The hyphenated LC-NMR technique for metabolomics profiling
Online hyphenation of LC-NMR has emerged as a powerful technique to provide crucial advantages over standalone NMR-based approaches, particularly when analyzing complex microbial extracts containing numerous metabolites exhibiting significant signal overlap.169,170 By efficiently separating compounds based on their physicochemical properties prior to NMR analysis, LC techniques significantly reduce the spectral complexity, enabling the identification and characterization of even low-abundance metabolites. Techniques such as WATERGATE171 and WET172 have been adapted for on-flow LC-NMR to minimize solvent interference, which otherwise interfere with the detection of low-abundance microbial NPs. In addition, LC-NMR can be further coupled with MS spectrometry and/or solid phase extraction (SPE) in a more sophisticated and automated approach (e.g. LC-SPE-NMR, LC-NMR-MS, and LC-SPE-NMR-MS),173 offering complementary information for enhanced metabolite identification and structure characterization. Online LC-SPE-NMR systems automate the preconcentration and cleanup steps, minimizing sample handling and loss. The target microbial NPs are selectively retained on an SPE cartridge, while interfering matrix components are washed away, resulting in a more concentrated and purified sample for NMR analysis. Different SPE sorbents allow the selective enrichment of compounds based on polarity or other characteristics.24 Furthermore, the coupling of LC-NMR with MS provides complementary information for comprehensive metabolite profiling. High-resolution MS offers accurate mass measurements and characteristic fragmentation patterns, aiding compound identification and dereplication.174 For example, Lin and colleagues developed a microscale LC-MS-NMR platform that combines nanoSplitter LC-MS and microdroplet NMR for identifying unknown compounds present at low concentrations within complex mixtures. The platform design allows highly sensitive MS analysis, while directing the majority of the LC effluent to NMR for structural characterization. By using a segmented flow microdroplet system coupled with high automation, interpretable 1H NMR spectra are obtained from analytes at the 200 ng level within 1 h acquisition. The utility of this platform is demonstrated through the identification of known and unknown metabolites from a bioactive cyanobacterial extract, showcasing its potential for streamlining the discovery of microbial NPs.175 In another separate report, the initial LC-NMR analysis of EtOAc extracts of several actinomycetes revealed an LC peak exhibiting multiple methyl doublet signals (δH 0.4–1.6) and α-methine signals (δH 4–5), suggesting a peptidic structure. The subsequent large-scale purification, monitored by LC-NMR, enabled the targeted discovery of tumescenamides A (53) and B (54) from Streptomyces tumescens YM23-260. In addition, Trimble et al. exemplified the utility of LC-NMR in resolving diastereomeric microbial NPs, namely (16R)-and (16S)-hydroxyroquefortine C (55 and 56), from Penicillium crustosum DAOM 215343.176
4. NMR-based metabolomics meets genomics
In the post-genomic era, genomic mining has become a routine approach for exploring microbial NPs. Notwithstanding, an effective platform is still indispensable to correlate the chemotype and genotype in the crude metabolomic extracts to efficiently avoid chemical redundancy and accelerate the targeted discovery process. The synergy between genomic and metabolomic approaches marks a cutting-edge development in the field of NP chemistry and sets the stage for a new phase of discovery.11,13,15,16 MS-based metabolomics has been well integrated with genomics, including the metabologenomics,177,178 peptidogenomics,179 glycogenomics,180 NPRminer,181 MetaMina,182 and NPLinker platforms.183 Also, a synergistic link exists between NMR-based metabolomics and genomics, and this interface can be analogously harnessed for bioprospecting.
4.1 General principle for paired genomics and NMR-based metabolomics
The underlying basic principle and/or workflow of paired NMR-based metabolomics and genomics for microbial NP discovery is illustrated in Fig. 5. In brief, advances in genomics, coupled with sophisticated bioinformatics tools, enable the rapid analysis of biosynthetic gene clusters (BGCs) within microbial genomes to discern their novelties through homology comparison.184–188 The accumulative biosynthetic knowledge allows the in silico structural prediction of chemical outputs of BGCs, including cursory chemical scaffolds and/or key functional groups. NMR spectroscopic data can be predicted with reasonable accuracy for bioinformatics-deduced theoretical chemical (sub)structures, including the chemical shifts and/or splitting patterns of each proton or carbon in 1H and 13C NMR spectra.189 Scrutiny of the corresponding NMR regions that potentially contain the predicted 1H and 13C resonances ensures the tracking of “biomarkers” arising from the BGC(s) of interest, achieving a direct correlation between the genotype and the chemotype of given BGCs. Despite the challenges in identifying all signals of a single molecule in a complex NMR fingerprint due to signal congestion, any distinguishable signals from (sub)structures predicted by BGCs can act as driven-peaks for the NMR-based matching of genotype and chemotype.
Fig. 5 Synergistic workflow of paired NMR-based metabolomics and genomics for the discovery of microbial NPs. The process begins with genomic analysis of microbial strains to identify BGCs. Bioinformatics analysis of the BGC of interest enable in silico structural prediction of cognate products. Then, the theoretically predicted NMR spectral data of metabolic products guides the targeted identification of characteristic “biomarkers” in the NMR profiling of crude extracts, establishing a genotype–chemotype correlation. Alternatively, the interpretation of experimental NMR profiling data allows the characterization of key structural fragments, which guide the assignment of responsible BGCs via retro-biosynthetic analysis. Then, sequence homology analysis of BGC and bioinformatics-aided structure prediction aid in the dereplication and/or structural novelty judgement of the final products represented by the characterized structural fragments.
Alternatively, the integration of these two disciplines can alternatively commence with 1D and 2D NMR analyses of microbial crude extracts to dereplicate known compounds or identify structural fragments of unknown compounds. The obtained structural signatures can be used as probes for the assignment of responsible BGCs on bacterial (meta)genomes, a process termed retro-biosynthetic analysis.190,191 Although NMR-based metabolomics alone may not be sufficient to dereplicate or assess the novelty of metabolites based on structural fragments, the suggested potential BGCs in microbial genomes reverts to the above-mentioned “gene-to-chemistry” strategy by directly comparing the BGC sequence homology to determine the structural novelty of the final products. Specifically, this synergistic approach facilitates more efficient dereplication and targeted discovery by enabling clearer spectroscopic signals and a more focused search for predicted metabolites.
4.1.1 Bioinformatics-aided structure elucidation and/or prediction. As illustrated in Fig. 5, bioinformatics-aided structure elucidation and/or prediction192 is a critical linchpin in paired genomics and NMR-based metabolomics strategies, bridging the gap between genomic information and metabolite characterization. In fact, bioinformatics-aided structure elucidation is complementary or even advantageous in some cases of spectroscopic approaches for structure elucidation. A prominent example is the discovery of bartolosides A–D, cyanobacterial glycolipids featuring aliphatic chains, whereby bioinformatics played a central role in their identification, otherwise formidable to accomplish solely by NMR analysis.193 Advanced bioinformatics algorithms and computational tools have been devised to forecast the chemical structures of BGCs solely from DNA sequences, and many excellent reviews have been published on this topic.184–188 This advancement is propelled by the expanding quantity of data on established biosynthetic elaboration.185,194 Particularly, the colinear nature of the biosynthetic machineries exemplified by type I modular polyketide synthases (PKS),195 modular non-ribosomal peptide synthetases (NRPS),196 and their hybrids,197 allows the direct prediction of the core structures of the resulting metabolites based on the domain architecture and substrate specificity of the encoding genes. For example, Terlouw and coworkers recently developed PARAS and PARASECT (https://paras.bioinformatics.nl), two novel machine learning-based algorithms for predicting the substrate specificity of adenylation (A) domains in NRPSs. By leveraging a large, high-quality training dataset and incorporating both sequence and, to a lesser extent, structural features of the A domains and their substrates, PARAS and PARASECT achieve significantly improved prediction accuracy compared to their counterpart computational tools such as NRPSPredictor2, AdenPredictor, and SANDPUMA.198 Ribosomally synthesized and post-translationally modified peptides (RiPPs) represent another class of microbial NPs amenable to genomics-aided structure elucidation. Biosynthesis involves ribosomal translation of a gene-encoded precursor peptide, followed by extensive enzymatic post-translational modifications (PTMs).199 Analysis of the precursor peptide sequence, combined with knowledge of conserved modification enzymes within the BGC, allows the prediction of core scaffold structures and potential post-translational modifications. Although these assembly lines readily enable genomics-aided structure prediction, some other common classes of BGCs, such as terpene synthases (TSs) and type II PKSs, are inherently more challenging to link to their products due to a lack of colinearity. In these cases, it is still not impossible to use specific enzymatic domains within BGCs to predict key structural features that can be paired with NMR-based metabolic profiling. For example, genes encoding post-modifications enzymes (e.g. O-methyltransferases and glycosyltransferase), uncommon building blocks and/or precursors (deoxyaminosugar cassette200) can provide valuable chemical handles for subsequent targeted characterization during NMR-based metabolic profiling at the mixture level. Specifically, if O-methyltransferase is present in the BGC of interest, a distinctive methoxy (δH 3.50–4.00) is anticipated in the NMR profiles.
Genomics-aided structure prediction has also found increasing applications in determining the absolute configurations of complex microbial NPs. By identifying stereospecific enzymatic domains within BGCs, such as ketoreductase (KR) domains in modular PKSs184,201,202 and epimerization (E) domains in NRPSs,203,204 researchers can infer the stereochemical outcome of key steps in the biosynthetic pathway. The knowledge of KR domains can lead to the understanding of the 3D chiral centers on PKS products or the specific stereochemistry of amino acids. This information, in conjunction with NMR spectroscopic data enables the more accurate and confident assignment of the absolute configuration of microbial NPs. For instance, Kim et al. showcased the power of integrating bioinformatic approaches and NMR analysis (coupling constants analysis and ROESY correlations) for the rapid and accurate stereochemical determination of neaumycin B, a structurally complex macrolide possessing 19 asymmetric centers.205
4.1.2 NMR profiling “readout” genomically predicted structures in mixtures. In a BGC of interest, a key challenge lies in rapidly identifying the corresponding product signals in the complex NMR profile of the microbial extract. The primary goal is to maximize the signal-to-noise ratio of the target compound relative to other interfering substances in the mixture. Firstly, methods to enhance the expression of the target (often silent) BGC are explored, including perturbations in the culture conditions and/or the application of synthetic biology tools (promoter engineering and heterologous expression).206 Further comparison of the treated experimental groups with control groups can facilitate the detection of NMR signals of interest. For example, if bioinformatics-aided structure prediction suggests a siderophore-encoding BGC in the host microbe, iron-deficient media can be employed to make the NMR signals of the target siderophore more prominent in complex biological matrices.207 Secondly, bioinformatics-aided structure prediction can be leveraged to rationally select extraction and separation methods that enrich the target products, while minimizing interfering components. For instance, if the BGC of interest encodes aromatic polyketides, ethyl acetate extraction is potentially superior to adsorption resin extraction, whereas NRPS BGCs encoding hydrophilic peptidic products are often missed by ethyl acetate extraction. A more specific example in this respect is the so-called “catch-release strategy” recently developed by the Hertweck group. They leveraged bioinformatics analysis to identify NRPS gene clusters encoding key enzymes critical for the biosynthesis of the extremely rare amino acid 3-furylalanine (Fua). The reactivity of furan rings and N-phenylmaleimide through Diels–Alder reaction enabled the direct enrichment of Fua-containing peptides from complex mixtures, thereby facilitating their detection and characterization.208 Thirdly, the integration of prefractionation strategies can further enhance the effectiveness of NMR-based metabolomics of genomics-predicted products. By reducing the complexity of crude extracts, the confidence and accuracy of correlating genomics-derived in silico structure predictions with NMR-based detection can be significantly improved. For instance, if the traced compounds are bioinformatically predicted to be relatively apolar polyenes, reverse-phase C18 separation of the metabolomes rinsed by 60% aqueous methanol can remove the common interferents such as the culture medium ingredients, and then enrich the components of interest, making it easier to find the sought-after signals in the NMR profiling of the simplified metabolomes. This strategy is particularly effective for the focused isolation of low-abundance microbial NPs. The illustrative examples can be found in Sections 4.3.
4.2 Genomic and NMR spectroscopic signature-based approach
A unique hybrid genomics and NMR-based metabolomics strategy termed “genomic and spectroscopic signature-based approach” has gained attention for streamlining the isolation and characterization of microbial NPs containing unique signature structural motifs. The general pipeline is summarized in Fig. 6. The creation of unique signature compounds often relies on conserved biosynthetic genes amenable to degenerating the primer design. High-throughput PCR screening based on the degenerate primers swiftly identifies bacterial strains harboring BGCs responsible for microbial NPs with unique structural motifs. The resulting PCR amplicons are sequenced and yield natural product sequence tags (NPSTs),209 which are analyzed phylogenetically to assess the novelty of BGCs. NPSTs situated in distinct clades from known references pinpoint strains possessing novel biosynthetic pathways.210,211 Genome sequencing and bioinformatics analysis of the prioritized strains confirm the novelty of targeted BGCs containing unique genetic probes. Next, the combined NMR-based metabolomics and genomics approach, as illustrated in Fig. 5, can be applied for targeted discovery. In brief, based on the chemical properties and NMR predictions of the signature motifs, appropriate NMR methods (such as 1H NMR, 31P-NMR, 1H–13C HSQC, 1H–15N HMBC, and 1H–15N HSQC-TOCSY) are selected for chemispecific detection from microbial extracts, followed by NMR-guided chromatographic separation. Identifying the spectroscopic signatures in microbial culture extracts purely using NMR experiments may be not feasible owing to the invariable complexity and signals congestion, especially for targeted BGCs expressed at a low level. This limitation can be overcome by isotope labelling of the precursors to enhance the detection sensitivity of signature structural motifs by NMR profiling, given that the key biosynthetic precursors for the unique signature structural motifs are already known. Overall, the combination of degenerate primer PCR screening, isotope labelling, and NMR metabolic profiling offers a general strategy for the identification of a diverse array of microbial NPs featuring unique signature structural motifs.
Fig. 6 Genomic and spectroscopic signature-based approach for the discovery of microbial NPs. Degenerate primer PCR screening, targeting conserved biosynthetic genes, identifies strains harboring BGCs for microbial NPs with signature structural unit. The sequencing of PCR amplicons and phylogenetic analysis assess the BGC novelty. The NMR-based metabolomic profiling tool is used for the straightforward identification of the microbial NPs derived from the BGCs of interest in the prioritized strains. Isotope-labelled precursor feeding can pinpoint the NMR signal(s) in highly complex matrices that are originated from the investigated BGCs, which are then used as probes for the NMR-guided separation of microbial NPs containing the signature structural unit.
4.3 Illustrative microbial NPs discovered by the joint strength of genomics and NMR-based metabolomics
In this section, we exemplify the proof of concept that the paired NMR-based metabolomics and genomics approaches, as discussed in Sections 4.2 and 4.3, can effectively accelerate the discovery rate of microbial NPs. The diversity in chemical structures and biosynthetic pathways elegantly supports the synergistic power of these two disciplines.
4.3.1 Type I modular PKS. Type I modular polyketide synthases (PKSs)195 represent a fascinating class of biosynthetic machineries responsible for the creation of structurally diverse polyketide NPs. The common structural features of type I PKS-derived polyketides include conjugated alkene double bonds, oxygenated carbons (hydroxyl, carbonyl, and epoxy), branched methyl groups, and cyclic structures (lactone or lactam ring). Given that this class of assembly lines is basically congruent with the colinearity rule, their metabolic products are well predictable by bioinformatics, and NMR-based metabolic profiling can effectively link the chemotype and genotype of these biosynthetic machineries at the crude extract stage. To exemplify this concept, our research group explored an approach by integrating genomics and 2D NMR-based metabolomics for the discovery of NPs encoded by type I modular PKSs from myxobacteria (Fig. 7). We began by sequencing the genome of our proprietary strain Archangium violaceum SDU8. Subsequent antiSMASH analysis revealed a unique trans-acyltransferase (trans-AT) PKS gene cluster (arc). Further analysis of the integral domains in the core PKS megasynthetases using the computational tool transATor212 suggested that arc potentially encodes an allenic polyketide. Driven by this prospect, we sought to rapidly target the arc-encoded product from the SDU8 crude extract using 2D NMR-based metabolomic profiling. The initial HSQC analysis using SMART 2.0,113 indicated that macrolides were the major components in the SDU8 crude extract. Subsequent inspection of the HMBC and COSY spectra further identified the typical structural fragments found in macrolides. Given that arc is the sole BGC capable of encoding a macrolide-like compound in the SDU8 genome, we could confidently correlate the arc BGC with the 2D NMR-derived structural features within the complex mixture. The scaled-up fermentation of SDU8 followed by NMR-guided separation facilitated the successful discovery of archangiumide (57),213 the founding member of allenic macrolide.
Fig. 7 Integrated genomics and NMR-based metabolomics workflow for the discovery of allenic macrolide. Bioinformatic analysis of a trans-AT PKS gene cluster arc in the A. violaceum SDU8 genome predicted its metabolic product as an allenic macrolide. Moreover, 2D NMR profiling of the SDU8 crude extract allowed the identification of the typical structural fragments of macrolides. This allowed the correlation of the genotype and chemotype of arc because it is the only BGC capable of biosynthesizing macrolide-class molecules in SDU8. Large-scale fermentation in tandem with NMR-guided separation streamlined the discovery of archangiumide (57).
The OH group employed the genomic and spectroscopic signature-based approach to discover macrolactams originating from the type I modular PKS machinery (Fig. 8).214 The biosynthesis of macrolactams is characterized by a unique protection and deprotection mechanism catalyzed by an adenylation enzyme (AD) and a proline aminopeptidase (PAP), respectively.215 Taking advantage of knowledge, the investigators performed PCR screening of the genomic DNA library using the degenerate primers specific to the AD and PAP genes, which allowed them to identify 43 macrolactam producers from a collection of 1188 bacterial strains. Phylogenetic analysis of the PAP amplicons enabled the reasonable prediction of the subtype and novelty of macrolactams produced by the 43 selected strains. To achieve the early identification of macrolactams in the microbial crude extracts without purification, 1H–15N HSQC-TOCSY NMR spectroscopic analysis was applied to display the expected 1H spin system on the amide nitrogen (δN 110−130 ppm), given that all three subtypes (α-methyl, α-alkyl, or β-methyl) of macrolactams commonly possess proton-proficient carbons (CH, CH2, CH3) connected to the amide NH. Considering the low natural abundance of 15N (0.36%), 15NH4Cl was fed to the cultures of the hit strains to strengthen the 15N detection, and thus facilitate the direct detection of the spectroscopic signatures by 1H–15N HSQC-TOCSY NMR experiments. Finally, this combined genomic and spectroscopic signature-based method resulted in the structural revision of an α-methyl macrolactam salinilactam (58), and the discovery of a new α-alkyl-type macrolactam muanlactam (59) and a new β-methyl macrolactam concolactam (60) with a [16,6,6]-tricyclic skeleton.
Fig. 8 Genomic and spectroscopic signature-based approach for the discovery of macrolactams. Degenerate primers targeting the conserved adenylation (AD) and proline aminopeptidase (PAP) genes were utilized to screen a bacterial strain library via PCR, identifying strains that harbor the biosynthetic pathway for macrolactams. The isotope labelling with 15NH4Cl followed by 1H–15N HSQC-TOCSY NMR profiling displayed the diagnostic 1H spin systems surrounding the amide nitrogen in three subtypes of macrolactams. This integrated approach led to the discovery of 58–60.
Likewise, the Oh group employed the same strategy to selectively discover oxazole-bearing natural products.216 A library of 1000 strains was screened via PCR using the degenerate primer specific to the key biosynthetic gene encoding oxazole cyclase. The unique one-bond coupling constants (1JCH values) and chemical shifts for the 1H and 13C atoms in oxazole provide diagnostic spectroscopic signatures in the 1H–13C HSQC NMR spectra. These signatures enable the selective and unequivocal detection of the terminal oxazole motif without the need for isotopic labeling or chromatographic purification of the strains selected by genomic screening. The application of this targeted method led to the discovery of five new compounds, including lenzioxazole (61), possessing an unprecedented cyclopentane, permafroxazole (62), bearing a tetraene conjugated with carboxylic acid, tenebriazine (63), incorporating two modified amino acids, and methyl-oxazolomycins A and B (64 and 65, respectively), expanding the chemical space of the terminal oxazole-bearing NP family.
4.3.2 Type II PKS. Although it is normally difficult to bioinformatically predict the full structures or backbones of the products from non-linear biosynthetic systems such as type II PKS, some chemical features are still predictable. Type II PKS orchestrate the aromatic polyketides that are rich in the resonances in the region of δH 6.5–7.5 of the 1H NMR spectrum, and oxygenation universally happens in this class of compounds, which gives characteristic HMBC correlations from δH 6.5–7.5 to δC 145–160. The existence of O-methyltransferase in the tailing enzyme region of a BGC most likely warrants a characteristic methyl singlet typical of δH 3.5–4.5 in the mid-field of the NMR spectrum. We exemplified the power of integrating genome mining with NMR-based metabolomic profiling for the rapid discovery of aromatic polyketides orchestrated by type II PKS (Fig. 9).97 The bioinformatic analysis of the genome of a soil-derived actinomycete Streptomyces sp. MBT76 identified a cryptic type II PKS gene cluster (qin) that was predicted to encode molecules featuring a N-methylated deoxyamino-sugar attached to an aromatic polyketide pyranonaphthoquinone backbone. After activating the cluster via overexpression of its pathway-specific regulator, comparative 1H NMR metabolic profiling revealed characteristic proton signals attributed to the predicted structural fragments in the mutant strain. This enabled rapid NMR-guided separation, leading to the efficient characterization of novel qinimycins (66–70).
Fig. 9 Paired genomics and NMR-based metabolomics strategy for the discovery of qinimycins.97 A cryptic type II PKS gene cluster (qin) contains genes predicted to encode molecules featuring deoxyamino-sugar forosamine (depicted in red), an aromatic polyketide pyranonaphthoquinone backbone (depicted in black), as well as a suite of genes associated with N-methylation. After the qin gene cluster was activated by the overexpression of its pathway-specific regulator, the comparative 1H NMR metabolic profiling enabled the detection of the characteristic proton signals assignable to the bioinformatics-predicted structural fragments in the mutant, as highlighted by the arrow. NMR-guided separation streamlined the characterization of qinimycins. The structural fragments of qinimycins are color-coded to match corresponding genes within qin.
4.3.3 Non-ribosomal peptides. Non-ribosomal peptides (NRPs) are a structurally diverse class of natural products synthesized by large multi-modular NRPSs.196 A notable characteristic of NRPS biosynthesis is the frequent incorporation of non-proteinogenic amino acids, whose synthesis is directed by conserved enzymatic domains often encoded within the NRPS BGC itself. Therefore, these conserved domains/enzymes can serve as a valuable genomic signature for the targeted discovery of novel NRPS-derived NPs containing these unique building blocks. By incorporating isotopically labeled precursors of these unusual amino acids into microbial cultures, researchers can track their incorporation into specific metabolites using NMR spectroscopy. This approach enhances the detection sensitivity for low-abundance non-ribosomal peptides and provides valuable insights into the biosynthetic origin and metabolic fate of the non-proteinogenic amino acids. For example, the Oh group applied the genomic and NMR spectroscopic signature-based method to discover piperazic acid (Piz)-containing depsipeptides.217 Biosynthetically, L-ornithine is converted to Piz by the piperazate synthases KtzI and KtzT,218 and then incorporated into an NRPS line by an adenylation domain.219 The ktzT gene was used as a genomic signature for the PCR screening of 2020 strains, and the resulting NPSTs from 62 strains were phylogenetically analyzed to assess the novelty of the Piz-containing peptides. Firstly, all the hit strains were subjected to LC/MS analysis, but it was ineffective to discern the Piz-bearing depsipeptides at the crude extract level. Subsequently, they developed a versatile method for the unequivocal detection of compounds incorporating various Piz-derived motifs with either “−N–NH−” or “−N–N”.219 The combination of isotopic labelling with cost-effective 15NH4Cl and a suite of 2D 1H–15N NMR experiments including 1H–15N HMBC, 1H–15N HSQC-TOCSY and 1H–15N HSQC provided a diagnostic spectroscopic signature at δN 72−83 ppm for Piz and δN 310−320 ppm for Dpz (dehydropiperazic acid). These consolidated genomic and spectroscopic signature-based analyses facilitated the characterization of the dimeric cyclohexapeptides chloptosin (71) and depsidomycin D (72) with two units of Dpz, polyoxyperuin seco acid (73), and 31-membered cyclic decapeptides lenziamides A and B (74 and 75). Similarly, the Andersen group identified the strain Streptomyces incarnatus NRRL 8089 as a potential producer of Piz-containing peptides through genome mining using ktzT as a marker. By feeding the strain with 15N2-L-Orn, they enhanced the detection of the desired molecules in 1H–15N HSQC-TOCSY NMR profiling, and successfully isolated the low-abundance incarnatapeptins A and B (76 and 77, respectively),220 and dentigerumycins F (78) and G (79).221 Huang et al.222 performed 1H NMR metabolic profiling of Streptomyces sp. S063, whereby unusual negative chemical shifts (δH −0.34 ppm) were revealed. NMR-guided fractionation coupled with HSQC analysis revealed that the fraction containing the probe signal was dominated by peptides featuring multiple N-methyl groups (NMe), which strongly suggested an NRPS biosynthetic origin. Meanwhile, genome analysis of Streptomyces sp. S063 identified six NRPS gene clusters, and one of them (cluster 20) contained multiple N-methyltransferase (nMT) domains, which correlated well with the NMR-based metabolic profiling data. Further bioinformatic analysis of cluster 20 revealed that its chemical output should be Piz-containing peptides, considering the co-existence of the ktzT and ktzI genes. The limited similarity (<48%) of cluster 20 to known Piz-containing peptides spurred large-scale fermentation for the targeted isolation of lenziamides D1 (80) and B1 (81).
The Gerwick group described the “genomisotopic approach” for the discovery of a non-ribosomal peptide. Bioinformatics analysis of the Pseudomonas fluorescens Pf-5 genome led to the identification of an orphan NRPS gene cluster. Subsequent substrate specificity analysis of the central NRPS megasynthetases confirmed that 15N-labeled leucine (Leu) is an optimal choice for isotope labeling of the encoded lipopeptide because four A domains accept Leu as building blocks, allowing a more robust NMR resonance signal from the resulting labeled product. Further 1H–15N HMBC NMR-based metabolic profiling of the isotope-enriched crude extract was instrumental in the discovery of orfamide A (82),223 a novel lipopeptide with surfactant activity.224 Very recently, Xu et al. combined genome mining and 2D NMR-based metabolomics for the targeted isolation of nocobactin-type lipopeptides (83–92) from phychrophillic Nocardia sp. L-016. Initially, small-scale cultures were grown, and the resulting ethyl acetate extract was subjected to HSQC and HMBC analysis to identify key structural moieties such as oxazoles and phenol units directly at the crude extract level. This metabolomic profiling informed subsequent genome mining, enabling the identification of the responsible BGCs encoding the observed structural features, exemplified by the nocobactin-like BGC (noc) harboring an oxazole-coding homolog (nocF). Further LC-MS/MS-based metabolomics coupled with molecular networking was performed to uncover novel analogues within the identified compound class. This combined approach leveraged NMR-derived structural insights to focus genomic searches, leading to the efficient discovery of novel natural products.
4.3.4 Terpenoids. Terpenoids are the largest and most structurally diverse family of NPs. Compared to fungi and plants, bacterial terpenoids make up a relatively small fraction (approximately 1%). Despite the small number of bacterial terpenoids that have been isolated, extensive genomic data indicates that nearly all bacteria possess significant potential for delivering a wide range of terpenoids.225 However, bacterial terpenoids are challenging to characterize due to their low yields and lack of chromophores that are important for detection using variable wavelength or diode array detection systems. Thus, to tackle this problem, Chen et al. recently developed a strategy by integrating genome mining with 1H NMR-based profiling for targeted discovery.226 The bioinformatic analysis of 20 actinomycetes prioritized the strain Crossiella cryophila for metabolomic investigation. Terpenoids are biosynthetically formed by the condensation of allylic dimethylallyl diphosphate (DMAPP) and homo-allylic isopentenyl diphosphate (IPP), resulting in final structures that always contain multiple methyl groups.227 The distinctive methyl protons that typically exhibit chemical shifts ranging from δH 0.60 to 1.85 ppm could serve as diagnostic signals for terpenoids mixed within intricate mixtures. After optimization of the culture media, 15 sesquiterpenoids with five distinct carbon skeletons, including nine new sesquiterpenoids (93–101), were isolated.
In another separate report, three unknown plant P450 genes were expressed in yeast to metabolize the terpenoids thalianol, arabidiol, and marneral. HSQC-based metabolomic profiling proved to be effective for analyzing the crude extracts, distinguishing trace triterpenoids that were indistinguishable in GC–MS. The side-by-side comparison of HSQC was particularly instrumental in evaluating the crude extracts from the cultures with and without a P450 gene candidate, enabling the straightforward identification of metabolite peaks of interest. Known compounds were readily identified by matching pairs of 1H–13C correlation with literature data, and new derivatives (102–104) featuring differential oxidation patterns were unambiguously elucidated by intensive 2D NMR analysis.228 This study demonstrated that the combination of genome mining, heterologous expression, and HSQC-based metabolomic profiling is an effective methodology to elucidate unknown metabolic pathways and rapidly identify their novel products.
4.3.5 RiPPs. Schmidt et al. combined NMR and metagenomic sequencing to characterize divamides (105), a type of structurally complex anti-HIV RiPPs encoded by symbiotic cyanobacteria in Didemnum molle tunicates.229 The limited resource supply impeded the full characterization of the divamides, but they could identify key structural fragments including the modified amino acids lanthionine, N-trimethylglutamate, and the partial peptide sequence “GTTR” based on the intensive NMR analysis. The obtained chemical signatures were used to probe the assembled metagenome of Didemnum molle tunicates, which resulted in the identification of div, a gene cluster containing all the requisite elements for the biosynthesis of the NMR-resolved substructures. Consequently, the identification of the correct BGC in the metagenomic sequencing boosted the 2D NMR analysis to complete the final structural determination of divamides, in a fashion of bioinformatics-aided structure elucidation.
5. Conclusion and perspectives
Metabolomics offers a powerful strategy to accelerate the discovery of microbial NPs by providing a comprehensive snapshot of the small molecules present in complex biological samples. MS and NMR are the two most commonly used technologies for metabolomics, primarily because of their powerful ability to detect individual metabolites in complex mixtures, while requiring little or no physical separation. However, NMR and MS have both strengths and weaknesses for metabolomics study.230 MS is more sensitive than NMR and gives better discriminatory signals due to its more common hyphenation to liquid chromatography, but MS is not suitable for molecules resistant to ionization. NMR can universally detect all compound classes present in a mixture, regardless of their ionization properties or chemical derivatization requirements. Furthermore, the noninvasive nature of NMR preserves the integrity of the sample for subsequent analyses, such as bioactivity testing or further structural characterization (e.g. crystallization). In terms of structure elucidation, NMR possesses the distinctive capability to investigate structural details at the atomic scale for NMR-active nuclei, including 1H, 13C, 15N, 11B, and 31P, enabling the structural determination of unknown compounds by providing detailed atom–atom linkages (e.g. C–H, C–C, C–O, C–N, C–P, and C–B). Moreover, NMR spectroscopy demonstrates a distinct advantage over MS in determining absolute configurations and distinguishing diastereomers.74,75 However, the low sensitivity and signal overlapping potentially make it challenging to detect important compounds present at lower concentrations by NMR-based metabolic profiling22 because they are likely to be overshadowed by predominant peaks.18 This problem is being actively addressed through various strategies. Prefractionation techniques, as discussed in Section 3, offer a direct approach by reducing the complexity of the mixture prior to NMR analysis. Sophisticated algorithms for automated spectral deconvolution can disentangle overlapping signals, while machine learning approaches can identify patterns and biomarkers even in highly congested regions of the spectrum (see Section 2.1.3). Techniques that enhance spectral resolution, such as high-field NMR and cryogenic probes (see Section 2.1.1), indirectly alleviate signal overlap by spreading signals further apart. Additionally, multi-dimensional NMR methods and isotopic labeling can also highlight the NMR signals of interest in complex mixtures.
The identification of unknown metabolites has been regarded as one of the major challenges in the metabolomics field, where NMR and MS alone are oftentimes not sufficient for the unambiguous identification of individual molecular species hidden in complex mixtures. Although NMR offers advantages for elucidating core structural frameworks of the predominant compounds, NMR metabolomics often struggles to resolve closely related analogues within complex mixtures. Conversely, MS, particularly when coupled with techniques such as GNPS molecular networking,231 excels at identifying and discriminating structural analogs through fragmentation analysis. Therefore, methods integrating both NMR and MS for the identification of metabolites in metabolomics have received particular attention, which heralds a promising future for navigating the complex chemical landscape of microbial metabolites with enhanced efficiency and precision.232 Brüschweiler and colleagues created tools such as SUMMIT233–235 and NMR/MS Translator,236 which integrate MS and NMR for the easy identification of unknowns in complex mixtures. The SUMMIT MS/NMR approach was used on an E. coli cell lysate, successfully identifying compounds such as N-acetylputrescine and spermidine.233 The NMR/MS Translator method begins with NMR spectra for structure identification, and then confirmed by MS analysis of potential ions, adducts, fragments, and isotope patterns.236 Ge et al. developed the Structure-Oriented Fractions Screening Platform (SFSP), a novel MS/NMR-hybrid metabolomics approach for accelerating natural product discovery.237 SFSP leverages machine learning algorithms to predict the 1H NMR chemical shift distributions from untargeted MS1 and 1H-PSYCHE NMR data, facilitating the functional-group-guided fractionation of complex mixtures. Unlike other metabolomics approaches, SFSP does not rely on pre-existing databases for compound identification, making it particularly useful for discovering novel compounds. This database-independent platform was successfully applied to the marine fungus Aspergillus sp. GE2-6, leading to the isolation and structural elucidation of 27 novel flavipidin and phenalenone derivatives. The integration of NMR and MS data within SFSP proved instrumental in the rapid dereplication of known compounds and the targeted identification of promising lead candidates based on characteristic NMR signals, thereby streamlining the natural product discovery workflow.
The effectiveness of genome mining is well-known,1,185,206 but it misses a considerable amount of undescribed chemistry due to the intricacy of NP biosynthesis, which happens in the constrained space of a cell, such as the involvement of non-enzymatic reactions238 and the crosstalk among multiple different BGCs.239,240 Besides, identifying phylogenetic uniqueness does not automatically ensure that a microorganism possesses the requisite biosynthetic pathways to produce new metabolites, or that these pathways are dormant. Thus, genomics needs to work in conjunction with metabolomics because the latter is directly geared towards the final stage of the metabolic process. In turn, metabolomics also reciprocally relies on genomics to guide targeted discovery and/or simplifies the structure elucidation in mixtures. Numerous studies have validated the working interface between genomics and MS-based metabolomics.15 Comparatively, the strategy of combining NMR-based metabolomics with genomics seriously lags behind. We reason that NMR-based metabolomics offers advantages over or at least complementary to MS-based metabolomics owing to the following: (1) NMR is more versatile in solving any characteristic structural fragments predicted from the biosynthetic cassettes of a given BGC, and thus more effectively bridges the gap between the chemical phenotypes and the genotypes. (2) More flexibly, by recognizing the signal distribution pattern in the NMR spectra, it can still match the fuzzy chemical categories predicted from novel BGCs, and thus the NMR-based metabolomics strategy is anticipated to find “unknown unknowns”.241,242 However, NMR-based metabolomics also has obvious limitations when combined with genomics, as follows: (1) the NMR-based metabolomics workflow relies on the expression level of BGCs of interest, otherwise the signals overlapping in the NMR spectra would be problematic, especially for minor components; (2) NMR-based metabolomics is less effective than MS-based approaches such as molecular networking in terms of uncovering structurally related congeners, especially for those produced in tiny amounts; and (3) NMR structural elucidation requires intensive expertise and experience, and thus is seemingly more difficult to be popularized than MS, especially for biological researchers, although the advancements in CASE84,85 for automated NMR interpretation can somehow mitigate this dilemma for non-spectroscopists. It is crucial to judiciously select between NMR and MS metabolic fingerprinting methods based on the specific types of BGCs. For example, MS-based metabolic profiling is more suitable for the RiPP class of compounds, whereas NMR-based metabolomics is more effective for detecting aromatic polyketides derived from type II PKS. Collectively, the pairwise complementarity of genomics, NMR-based metabolomics and MS-based metabolomics [“genomics-MS” and “genomics-NMR”, “NMR-MS”] implicates the feasibility of integrating three methodologies as a new platform to provide a new impetus for the discovery of novel microbial NPs, given that they offer different strengths but also display drawbacks.
Finally, AI is poised to revolutionize the discovery of microbial NPs by accelerating and refining various aspects of the process. From genome mining and BGCs prediction to structure elucidation and bioactivity prediction, AI algorithms offer the potential to overcome longstanding bottlenecks in this field.243–245 In the context of NMR-based metabolomics, AI can significantly enhance data processing, spectral analysis, and compound identification. For example, deep learning models can be trained to automate peak picking, chemical shift prediction, and spectral deconvolution, improving the speed and accuracy of NMR data interpretation, as discussed in Section 2. Furthermore, AI can facilitate the correlation of NMR spectral features with bioactivity data, enabling the rapid identification of lead compounds and the design of more targeted discovery campaigns. By integrating AI with NMR-based metabolomics workflows, researchers can unlock new avenues for microbial NP discovery and gain a deeper understanding of microbial chemistry.51,101,246,247 It is conceivable that the continued development of more AI technologies can revolutionize aspects of microbial NP discovery, especially in conjunction with orthogonal information derived from genomics and MS-based metabolomics. Furthermore, there is a growing emphasis from the NP community on the sharing and dissemination of NMR raw data.248 This trend is crucial for supplying ample datasets for the optimization of deep machine learning models, thereby reinforcing the application of AI in the paradigm of NMR-based metabolomics and enhancing its analytical capabilities. Also, it is urgently necessary to develop a reliable protocol with initiatives of integrating the existing algorithms for genomics188,192 and NMR-based compound annotation in mixtures. A quick way to boost the strength is probably to emulate the successful MS-mediated linking of genomics and metabolomics.177–183 For instance, the binary correlation between the biosynthetic gene cluster families (GCFs) and corresponding NPs relies on similarity scoring across data sets,177 which can be potentially referenced by NMR-based metabolomics in the differential analysis of a bunch of genetically-associated bacterial strains such as actinomycetes and myxobacteria. In the future, huge opportunities exist for the development of sophisticated computational tools to integrate genomics, MS, and NMR datasets as a whole and use them synergistically for the bioprospecting of new bioactive compounds as well as associated enzymology in an automated and high-throughput manner, which will require extensive interdisciplinary knowledge and cooperation of researchers from chemistry, biology, and cheminformatics.
6. Conflicts of interest
There are no conflicts to declare.
7. Acknowledgements
This work was financially supported by the National Key Research and Development Programs of China (No. 2024YFA0917100), the National Natural Science Foundation of China (NSFC) (No. 32222003 and 82304353), the Fundamental Research Funds of Shandong University (No. 2023QNTD001, and ts20230201), Postdoctoral Fellowship Program of CPSF (No. GZC20231463), the Natural Science Foundation of Shandong Province (No. ZR2024QB114), and the Postdoctoral Innovation Project of Shandong Province (No. SDCX-ZG-202303002).
8. References
A. G. Atanasov, S. B. Zotchev and V. M. Dirsch, Nat. Rev. Drug Discovery, 2021, 20, 200–216 CrossRefCASPubMed.
A. Gavriilidou, S. A. Kautsar, N. Zaburannyi, D. Krug, R. Müller, M. H. Medema and N. Ziemert, Nat. Microbiol., 2022, 7, 726–735 CrossRefCASPubMed.
C. R. Pye, M. J. Bertin, R. S. Lokey, W. H. Gerwick and R. G. Linington, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 5601–5606 CrossRefCASPubMed.
C. Wu, H. K. Kim, G. P. Van Wezel and Y. H. Choi, Drug Discov. Today Technol., 2015, 13, 11–17 CrossRefPubMed.
D. Krug and R. Müller, Nat. Prod. Rep., 2014, 31, 768–783 RSC.
B. C. Covington, J. A. McLean and B. O. Bachmann, Nat. Prod. Rep., 2017, 34, 6–24 RSC.
H. K. Kim, Y. H. Choi and R. Verpoorte, Trends Biotechnol., 2011, 29, 267–275 CrossRefCASPubMed.
H. K. Kim, Y. H. Choi and R. Verpoorte, Nat. Protoc., 2010, 5, 536–549 CrossRefCASPubMed.
B. C. Covington and M. R. Seyedsayamdost, Nat. Prod. Rep., 2025 10.1039/d4np00026a.
A. Bauermeister, H. Mannochio-Russo, L. V. Costa-Lotufo, A. K. Jarmusch and P. C. Dorrestein, Nat. Rev. Microbiol., 2022, 20, 143–160 CrossRefCASPubMed.
N. E. Avalon, A. E. Murray and B. J. Baker, Anal. Chem., 2022, 94, 11959–11966 CrossRefCASPubMed.
M. A. Schorn, S. Verhoeven, L. Ridder, F. Huber, D. D. Acharya, A. A. Aksenov, G. Aleti, J. A. Moghaddam, A. T. Aron, S. Aziz, A. Bauermeister, K. D. Bauman, M. Baunach, C. Beemelmanns, J. M. Beman, M. V. Berlanga-Clavero, A. A. Blacutt, H. B. Bode, A. Boullie, A. Brejnrod, T. S. Bugni, A. Calteau, L. Cao, V. J. Carrión, R. Castelo-Branco, S. Chanana, A. B. Chase, M. G. Chevrette, L. V. Costa-Lotufo, J. M. Crawford, C. R. Currie, B. Cuypers, T. Dang, T. de Rond, A. M. Demko, E. Dittmann, C. Du, C. Drozd, J. C. Dujardin, R. J. Dutton, A. Edlund, D. P. Fewer, N. Garg, J. M. Gauglitz, E. C. Gentry, L. Gerwick, E. Glukhov, H. Gross, M. Gugger, D. G. Guillén Matus, E. J. N. Helfrich, B. F. Hempel, J. S. Hur, M. Iorio, P. R. Jensen, K. Bin Kang, L. Kaysser, N. L. Kelleher, C. S. Kim, K. H. Kim, I. Koester, G. M. König, T. Leao, S. R. Lee, Y. Y. Lee, X. Li, J. C. Little, K. N. Maloney, D. Männle, C. Martin H, A. C. McAvoy, W. W. Metcalf, H. Mohimani, C. Molina-Santiago, B. S. Moore, M. W. Mullowney, M. Muskat, L. F. Nothias, E. C. O'Neill, E. I. Parkinson, D. Petras, J. Piel, E. C. Pierce, K. Pires, R. Reher, D. Romero, M. C. Roper, M. Rust, H. Saad, C. Saenz, L. M. Sanchez, S. J. Sørensen, M. Sosio, R. D. Süssmuth, D. Sweeney, K. Tahlan, R. J. Thomson, N. J. Tobias, A. E. Trindade-Silva, G. P. van Wezel, M. Wang, K. C. Weldon, F. Zhang, N. Ziemert, K. R. Duncan, M. Crüsemann, S. Rogers, P. C. Dorrestein, M. H. Medema and J. J. J. van der Hooft, Nat. Chem. Biol., 2021, 17, 363–368 CrossRefCASPubMed.
F. Panter, C. D. Bader and R. Müller, Chem. Sci., 2021, 12, 5994–6010 RSC.
L. K. Caesar, R. Montaser, N. P. Keller and N. L. Kelleher, Nat. Prod. Rep., 2021, 38, 2041–2065 RSC.
J. J. J. van der Hooft, H. Mohimani, A. Bauermeister, P. C. Dorrestein, K. R. Duncan and M. H. Medema, Chem. Soc. Rev., 2020, 49, 3297–3314 RSC.
S. Soldatou, G. H. Eldjarn, A. Huerta-Uribe, S. Rogers and K. R. Duncan, FEMS Microbiol. Lett., 2019, 366, 1–8 CrossRefPubMed.
G. A. Nagana Gowda and D. Raftery, Anal. Chem., 2017, 89, 490–510 CrossRefCASPubMed.
J. L. Markley, R. Brüschweiler, A. S. Edison, H. R. Eghbalnia, R. Powers, D. Raftery and D. S. Wishart, Curr. Opin. Biotechnol., 2017, 43, 34–40 CrossRefCASPubMed.
C. K. Larive, G. A. Barding and M. M. Dinges, Anal. Chem., 2015, 87, 133–146 CrossRefCASPubMed.
O. Beckonert, H. C. Keun, T. M. D. Ebbels, J. Bundy, E. Holmes, J. C. Lindon and J. K. Nicholson, Nat. Protoc., 2007, 2, 2692–2703 CrossRefCASPubMed.
R. Powers, E. R. Andersson, A. L. Bayless, R. B. Brua, M. C. Chang, L. L. Cheng, C. S. Clendinen, D. Cochran, V. Copié, J. R. Cort, A. A. Crook, H. R. Eghbalnia, A. Giacalone, G. J. Gouveia, J. C. Hoch, M. J. Jeppesen, A. S. Maroli, M. E. Merritt, W. Pathmasiri, H. E. Roth, A. Rushin, I. T. Sakallioglu, S. Sarma, T. B. Schock, L. W. Sumner, P. Takis, M. Uchimiya and D. S. Wishart, TrAC, Trends Anal. Chem., 2024, 171, 117478 CrossRefCASPubMed.
I. Garcia-Perez, J. M. Posma, J. I. Serrano-Contreras, C. L. Boulangé, Q. Chan, G. Frost, J. Stamler, P. Elliott, J. C. Lindon, E. Holmes and J. K. Nicholson, Nat. Protoc., 2020, 15, 2538–2567 CrossRefCASPubMed.
P. J. Rutledge and G. L. Challis, Nat. Rev. Microbiol., 2015, 13, 509–523 CrossRefCASPubMed.
M. Månsson, R. K. Phipps, L. Gram, M. H. G. Munro, T. O. Larsen and K. F. Nielsen, J. Nat. Prod., 2010, 73, 1126–1132 CrossRefPubMed.
C. D. Bader, M. Neuber, F. Panter, D. Krug and R. Müller, Anal. Chem., 2020, 92, 15403–15411 CrossRefCASPubMed.
Z. Gan, I. Hung, X. Wang, J. Paulino, G. Wu, I. M. Litvak, P. L. Gor’kov, W. W. Brey, P. Lendi, J. L. Schiano, M. D. Bird, I. R. Dixon, J. Toth, G. S. Boebinger and T. A. Cross, J. Magn. Reson., 2017, 284, 125–136 CrossRefCASPubMed.
J. N. Thomas, V. Ramaswamy, I. M. Litvak, T. L. Johnston, A. S. Edison and W. W. Brey, IEEE Trans. Appl. Supercond., 2021, 31, 13–16 Search PubMed.
B. Wouters, P. Miggiels, R. Bezemer, E. A. W. Van Der Cruijsen, E. Van Leeuwen, J. Gauvin, K. Houben, K. Babu Sai Sankar Gupta, P. Zuijdwijk, A. Harms, A. Carvalho De Souza and T. Hankemeier, Anal. Chem., 2022, 94, 15350–15358 CrossRefCASPubMed.
R. W. Adams, C. M. Holroyd, J. A. Aguilar, M. Nilsson and G. A. Morris, Chem. Commun., 2013, 49, 358–360 RSC.
Y. Nishiyama, G. Hou, V. Agarwal, Y. Su and A. Ramamoorthy, Chem. Rev., 2023, 123, 918–988 CrossRefCASPubMed.
D. Li, A. L. Hansen, L. Bruschweiler-Li and R. Brüschweiler, Chem.–Eur. J., 2018, 24, 11535–11544 CrossRefCASPubMed.
S. Robson, H. Arthanari, S. G. Hyberts and G. Wagner, in Methods in Enzymology, Elsevier Inc, 1st edn, 2019, vol. 614, pp. 263–29190 Search PubMed.
M. Jaeger and R. L. E. G. Aspers, in Annual Reports on NMR Spectroscopy, Elsevier Ltd, 1st edn, 2014, vol. 83, pp. 271–34990 Search PubMed.
A. L. Hansen and R. Brüschweiler, Angew. Chem., Int. Ed., 2016, 55, 14169–14172 CrossRefCASPubMed.
Ē. Kupče and T. D. W. Claridge, Angew. Chem., Int. Ed., 2017, 56, 11779–11783 CrossRefPubMed.
A. L. Hansen, R. Kupče, D. W. Li, L. Bruschweiler-Li, C. Wang and R. Brüschweiler, Anal. Chem., 2021, 93, 6112–6119 CrossRefCASPubMed.
M. Sawall, E. von Harbou, A. Moog, R. Behrens, H. Schröder, J. Simoneau, E. Steimers and K. Neymeyr, J. Magn. Reson., 2018, 289, 132–141 CrossRefCASPubMed.
L. Chen, Z. Weng, L. Y. Goh and M. Garland, J. Magn. Reson., 2002, 158, 164–168 CrossRefCAS.
H. de Brouwer, J. Magn. Reson., 2009, 201, 230–238 CrossRefCASPubMed.
V. Zorin, M. A. Bernstein and C. Cobas, Magn. Reson. Chem., 2017, 55, 738–746 CrossRefCASPubMed.
S. Golotvin and A. Williams, J. Magn. Reson., 2000, 146, 122–125 CrossRefCASPubMed.
K. C. Wang, S. Y. Wang, C. H. Kuo and Y. J. Tseng, Anal. Chem., 2013, 85, 1231–1239 CrossRefCASPubMed.
R. K. Harris, E. D. Becker, S. M. Cabral De Menezes, P. Granger, R. E. Hoffman and K. W. Zilm, Pure Appl. Chem., 2008, 80, 59–84 CrossRefCAS.
R. K. Harris, E. D. Becker, S. M. C. De Menezes, R. Goodfellow and P. Granger, Solid State Nucl. Magn. Reson., 2002, 22, 458–483 CrossRefCASPubMed.
H. Du, X. Gu, J. Chen, C. Bai, X. Duan and K. Hu, Anal. Chem., 2023, 95, 3195–3203 CrossRefCASPubMed.
J. J. Helmus and C. P. Jaroniec, J. Biomol. NMR, 2013, 55, 355–367 CrossRefCASPubMed.
M. Rahimi, Y. Lee, J. L. Markley and W. Lee, J. Magn. Reson., 2021, 328, 106995 CrossRefCASPubMed.
M. Priessner, Nat. Rev. Chem., 2023, 7, 672 CrossRefPubMed.
L. H. E. Wieske, S. Peintner and M. Erdélyi, Nat. Rev. Chem., 2023, 7, 511–524 CrossRefCASPubMed.
D. Chen, Z. Wang, D. Guo, V. Orekhov and X. Qu, Chem.–Eur. J., 2020, 26, 10391–10401 CrossRefCASPubMed.
N. Schmid, S. Bruderer, F. Paruzzo, G. Fischetti, G. Toscano, D. Graf, M. Fey, A. Henrici, V. Ziebart, B. Heitmann, H. Grabner, J. D. Wegner, R. K. O. Sigel and D. Wilhelm, J. Magn. Reson., 2023, 347, 107357 CrossRefCASPubMed.
D. W. Li, A. L. Hansen, C. Yuan, L. Bruschweiler-Li and R. Brüschweiler, Nat. Commun., 2021, 12, 5229 CrossRefCASPubMed.
D. W. Li, A. Leggett, L. Bruschweiler-Li and R. Brüschweiler, Anal. Chem., 2022, 94, 8674–8682 CrossRefCASPubMed.
D. W. Li, L. Bruschweiler-Li, A. L. Hansen and R. Brüschweiler, Magn. Reson., 2023, 4, 19–26 CrossRefCASPubMed.
M. Meunier, A. Schinkovitz and S. Derbré, Nat. Prod. Rep., 2024, 41, 1766–1786 RSC.
H. Parastar and R. Tauler, Angew Chem. Int. Ed. Engl., 2022, 61, e201801134 CrossRefCASPubMed.
S. A. A. Sousa, A. Magalhães and M. M. C. Ferreira, Chemom. Intell. Lab. Syst., 2013, 122, 93–102 CrossRefCAS.
T. De Meyer, D. Sinnaeve, B. Van Gasse, E. Tsiporkova, E. R. Rietzschel, M. L. De Buyzere, T. C. Gillebert, S. Bekaert, J. C. Martins and W. Van Criekinge, Anal. Chem., 2008, 80, 3783–3790 CrossRefCASPubMed.
G. Kleks, D. C. Holland, J. Porter and A. R. Carroll, Chem. Sci., 2021, 12, 10930–10943 RSC.
S. P. Gaudêncio and F. Pereira, Nat. Prod. Rep., 2015, 32, 779–810 RSC.
C. L. Zani and A. R. Carroll, J. Nat. Prod., 2017, 80, 1758–1766 CrossRefCASPubMed.
J. J. Ellinger, R. A. Chylla, E. L. Ulrich and J. L. Markley, Curr. Metabolomics, 2012, 1, 28–40 Search PubMed.
J. C. Hoch, K. Baskaran, H. Burr, J. Chin, H. R. Eghbalnia, T. Fujiwara, M. R. Gryk, T. Iwata, C. Kojima, G. Kurisu, D. Maziuk, Y. Miyanoiri, J. R. Wedell, C. Wilburn, H. Yao and M. Yokochi, Nucleic Acids Res., 2023, 51, D368–D376 CrossRefCASPubMed.
D. S. Wishart, Y. D. Feunang, A. Marcu, A. C. Guo, K. Liang, R. Vázquez-Fresno, T. Sajed, D. Johnson, C. Li, N. Karu, Z. Sayeeda, E. Lo, N. Assempour, M. Berjanskii, S. Singhal, D. Arndt, Y. Liang, H. Badran, J. Grant, A. Serra-Cayuela, Y. Liu, R. Mandal, V. Neveu, A. Pon, C. Knox, M. Wilson, C. Manach and A. Scalbert, Nucleic Acids Res., 2018, 46, D608–D617 CrossRefCASPubMed.
K. Bingol, L. Bruschweiler-Li, D. W. Li and R. Bruüschweiler, Anal. Chem., 2014, 86, 5494–5501 CrossRefCASPubMed.
K. Bingol, F. Zhang, L. Bruschweiler-Li and R. Brüschweiler, Anal. Chem., 2012, 84, 9395–9401 CrossRefCASPubMed.
D.-G. Wang, C.-Y. Wang, J.-Q. Hu, J.-J. Wang, W.-C. Liu, W.-J. Zhang, X.-R. Du, H. Wang, L.-L. Zhu, H.-Y. Sui, Y.-Z. Li and C. Wu, Anal. Chem., 2023, 95, 5256–5266 CrossRefCASPubMed.
A. F. A. Moumbock, M. Gao, A. Qaseem, J. Li, P. A. Kirchner, B. Ndingkokhar, B. D. Bekono, C. V. Simoben, S. B. Babiaka, Y. I. Malange, F. Sauter, P. Zierep, F. Ntie-Kang and S. Günther, Nucleic Acids Res., 2021, 49, D600–D604 CrossRefCASPubMed.
M. R. Jones, E. Pinto, M. A. Torres, F. Dörr, H. Mazur-Marzec, K. Szubert, L. Tartaglione, C. Dell'Aversano, C. O. Miles, D. G. Beach, P. McCarron, K. Sivonen, D. P. Fewer, J. Jokela and E. M. L. Janssen, Water Res., 2021, 196, 117017 CrossRefCASPubMed.
J. A. Van Santen, G. Jacob, A. L. Singh, V. Aniebok, M. J. Balunas, D. Bunsko, F. C. Neto, L. Castaño-Espriu, C. Chang, T. N. Clark, J. L. Cleary Little, D. A. Delgadillo, P. C. Dorrestein, K. R. Duncan, J. M. Egan, M. M. Galey, F. P. J. Haeckl, A. Hua, A. H. Hughes, D. Iskakova, A. Khadilkar, J. H. Lee, S. Lee, N. Legrow, D. Y. Liu, J. M. Macho, C. S. McCaughey, M. H. Medema, R. P. Neupane, T. J. O'Donnell, J. S. Paula, L. M. Sanchez, A. F. Shaikh, S. Soldatou, B. R. Terlouw, T. A. Tran, M. Valentine, J. J. J. Van Der Hooft, D. A. Vo, M. Wang, D. Wilson, K. E. Zink and R. G. Linington, ACS Cent. Sci., 2019, 5, 1824–1833 CrossRefCASPubMed.
M. Sorokina and C. Steinbeck, J. Cheminf., 2020, 12, 1–51 Search PubMed.
G. Bifulco, P. Dambruoso, L. Gomez-Paloma and R. Riccio, Chem. Rev., 2007, 107, 3744–3779 CrossRefCASPubMed.
S. G. Smith and J. M. Goodman, J. Am. Chem. Soc., 2010, 132, 12946–12959 CrossRefCASPubMed.
A. Mándi and T. Kurtán, Nat. Prod. Rep., 2019, 36, 889–918 RSC.
T. R. Hoye, C. S. Jeffrey and F. Shao, Nat. Protoc., 2007, 2, 2451–2458 CrossRefCASPubMed.
R. Bhushan and H. Brückner, J. Chromatogr. B:Anal. Technol. Biomed. Life Sci., 2011, 879, 3148–3161 CrossRefCASPubMed.
L. Gross, F. Mohn, N. Moll, G. Meyer, R. Ebel, W. M. Abdel-Mageed and M. Jaspars, Nat. Chem., 2010, 2, 821–825 CrossRefCASPubMed.
W. He, Y. Yu, K. Iizuka, H. Takezawa and M. Fujita, Nat. Chem., 2025, 17, 653–662 CrossRefCASPubMed.
Y. Inokuma, S. Yoshioka, J. Ariyoshi, T. Arai, Y. Hitora, K. Takada, S. Matsunaga, K. Rissanen and M. Fujita, Nature, 2013, 495, 461–466 CrossRefCASPubMed.
J. G. Song, J. Zheng, R. J. Wei, Y. L. Huang, J. Jiang, G. H. Ning, Y. Wang, W. Lu, W. C. Ye and D. Li, Chem, 2024, 10, 924–937 CAS.
C. G. Jones, M. W. Martynowycz, J. Hattne, T. J. Fulton, B. M. Stoltz, J. A. Rodriguez, H. M. Nelson and T. Gonen, ACS Cent. Sci., 2018, 4, 1587–1592 CrossRefCASPubMed.
D. C. Burns, E. P. Mazzola and W. F. Reynolds, Nat. Prod. Rep., 2019, 36, 919–933 RSC.
M. Elyashberg and D. Argyropoulos, Magn. Reson. Chem., 2021, 59, 669–690 CrossRefCASPubMed.
P. Kessler and M. Godejohann, Magn. Reson. Chem., 2018, 56, 480–492 CrossRefCASPubMed.
E. Troche-Pesqueira, C. Anklin, R. R. Gil and A. Navarro-Vázquez, Angew. Chem., Int. Ed., 2017, 56, 3660–3664 CrossRefCASPubMed.
F. Hallwass, M. Schmidt, H. Sun, A. Mazur, G. Kummerlöwe, B. Luy, A. Navarro-Vázquez, C. Griesinger and U. M. Reinscheid, Angew. Chem., Int. Ed., 2011, 50, 9487–9490 CrossRefCASPubMed.
A. V. Buevich and M. E. Elyashberg, Magn. Reson. Chem., 2020, 58, 594–606 CrossRefCASPubMed.
I. M. Novitskiy and A. G. Kutateladze, J. Org. Chem., 2022, 87, 4818–4828 CrossRefCASPubMed.
I. M. Novitskiy and A. G. Kutateladze, Nat. Prod. Rep., 2022, 39, 2003–2007 RSC.
I. Cortes, C. Cuadrado, A. Hernández Daranas and A. M. Sarotti, Front. Nat. Prod. Res., 2023, 2, 1–11 Search PubMed.
W. J. Ai, J. Li, D. Cao, S. Liu, Y. Y. Yuan, Y. Li, G. S. Tan, K. P. Xu, X. Yu, F. Kang, Z. X. Zou and W. X. Wang, J. Nat. Prod., 2024, 87, 743–752 CrossRefCASPubMed.
Q. Zhao, Y. Liu, H. Ma, Y. Qiao, J. Chao, X. Hou, Y. Wang and Y. Wang, Anal. Chim. Acta, 2020, 1110, 131–140 CrossRefCASPubMed.
C. Wu, B. Zacchetti, A. F. J. Ram, G. P. Van Wezel, D. Claessen and Y. H. Choi, Sci. Rep., 2015, 5, 1–10 Search PubMed.
C. Wu, C. Du, J. Gubbens, Y. H. Choi and G. P. Van Wezel, J. Nat. Prod., 2015, 78, 2355–2363 CrossRefCASPubMed.
C. Wu, C. Du, K. Ichinose, Y. H. Choi and G. P. Van Wezel, J. Nat. Prod., 2017, 80, 269–277 CrossRefCASPubMed.
C. Wu, H. Zhu, G. P. van Wezel and Y. H. Choi, Metabolomics, 2016, 12(12) Search PubMed.
J. Gubbens, H. Zhu, G. Girard, L. Song, B. I. Florea, P. Aston, K. Ichinose, D. V. Filippov, Y. H. Choi, H. S. Overkleeft, G. L. Challis and G. P. Van Wezel, Chem. Biol., 2014, 21, 707–718 CrossRefCASPubMed.
C. Wu, K. Ichinose, Y. H. Choi and G. P. van Wezel, ChemBioChem, 2017, 18, 1428–1434 CrossRefCASPubMed.
K. Wu, J. Luo, Q. Zeng, X. Dong, J. Chen, C. Zhan, Z. Chen and Y. Lin, Anal. Chem., 2021, 93, 1377–1382 CrossRefCASPubMed.
J. Hubert, J. M. Nuzillard, S. Purson, M. Hamzaoui, N. Borie, R. Reynaud and J. H. Renault, Anal. Chem., 2014, 86, 2955–2962 CrossRefCASPubMed.
Z. Yang, Y. Wu, H. Zhou, X. Cao, X. Jiang, K. Wang and S. Wu, RSC Adv., 2015, 5, 77553–77564 RSC.
A. Bakiri, J. Hubert, R. Reynaud, S. Lanthony, D. Harakat, J. H. Renault and J. M. Nuzillard, J. Nat. Prod., 2017, 80, 1387–1396 CrossRefCASPubMed.
C. S. Clendinen, B. Lee-McMullen, C. M. Williams, G. S. Stupp, K. Vandenborne, D. A. Hahn, G. A. Walter and A. S. Edison, Anal. Chem., 2014, 86, 9242–9250 CrossRefCASPubMed.
A. Bruguière, S. Derbré, D. Breárd, F. Tomi, J. M. Nuzillard and P. Richomme, Planta Med., 2021, 87, 1061–1068 CrossRefPubMed.
A. Bruguière, S. Derbré, J. Dietsch, J. Leguy, V. Rahier, Q. Pottier, D. Bréard, S. Suor-Cherer, G. Viault, A. M. Le Ray, F. Saubion and P. Richomme, Anal. Chem., 2020, 92, 8793–8801 CrossRefPubMed.
T. Yin, Y. Yu, Q. Liu, G. Zhu, L. Bai, W. Zhang and Z. Jiang, J. Food Compos. Anal., 2023, 115, 1–8 Search PubMed.
B. M. Duggan, R. Cullum, W. Fenical, L. A. Amador, A. D. Rodríguez and J. J. La Clair, Angew. Chem., Int. Ed., 2020, 59, 1144–1148 CrossRefCASPubMed.
J. L. Markley, E. L. Ulrich, H. M. Berman, K. Henrick, H. Nakamura and H. Akutsu, J. Biomol. NMR, 2008, 40, 153–155 CrossRefCASPubMed.
M. K. Renner, Y. C. Shen, X. C. Cheng, P. R. Jensen, W. Frankmoelle, C. A. Kauffman, W. Fenical, E. Lobkovsky and J. Clardy, J. Am. Chem. Soc., 1999, 121, 11273–11276 CrossRefCAS.
C. Zhang, Y. Idelbayev, N. Roberts, Y. Tao, Y. Nannapaneni, B. M. Duggan, J. Min, E. C. Lin, E. C. Gerwick, G. W. Cottrell and W. H. Gerwick, Sci. Rep., 2017, 7, 1–17 CrossRefPubMed.
R. Reher, H. W. Kim, C. Zhang, H. H. Mao, M. Wang, L. F. Nothias, A. M. Caraballo-Rodriguez, E. Glukhov, B. Teke, T. Leao, K. L. Alexander, B. M. Duggan, E. L. Van Everbroeck, P. C. Dorrestein, G. W. Cottrell and W. H. Gerwick, J. Am. Chem. Soc., 2020, 142, 4114–4120 CrossRefCASPubMed.
H. W. Kim, C. Zhang, R. Reher, M. Wang, K. L. Alexander, L. F. Nothias, Y. K. Han, H. Shin, K. Y. Lee, K. H. Lee, M. J. Kim, P. C. Dorrestein, W. H. Gerwick and G. W. Cottrell, J. Cheminf., 2023, 15, 71 Search PubMed.
A. Bakiri, J. Hubert, R. Reynaud, C. Lambert, A. Martinez, J. H. Renault and J. M. Nuzillard, J. Chem. Inf. Model., 2018, 58, 262–270 CrossRefCASPubMed.
T. P. Yin, Y. Yu, Q. H. Liu, M. Y. Zhou, G. Y. Zhu, L. P. Bai, W. Zhang and Z. H. Jiang, Chin. J. Chem., 2022, 40, 2169–2178 CrossRefCAS.
F. Bhinderwala, T. Vu, T. G. Smith, J. Kosacki, D. D. Marshall, Y. Xu, M. Morton and R. Powers, Anal. Chem., 2022, 94, 16308–16318 CrossRefCASPubMed.
K. Bingol and R. Bruschweiler, Anal. Chem., 2011, 83, 7412–7417 CrossRefCASPubMed.
J. M. Egan, J. A. van Santen, D. Y. Liu and R. G. Linington, J. Nat. Prod., 2021, 84, 1044–1055 CrossRefCASPubMed.
L. Buedenbender, L. J. Habener, T. Grkovic, D. I. Kurtböke, S. Duffy, V. M. Avery and A. R. Carroll, J. Nat. Prod., 2018, 81, 957–965 CrossRefCASPubMed.
L. Buedenbender, L. P. Robertson, L. Lucantoni, V. M. Avery, D. I. K. Ke and A. R. Carroll, Mar. Drugs, 2018, 16, 189–199 CrossRefPubMed.
F. Zhang and R. Brüschweiler, Angew. Chem., Int. Ed., 2007, 46, 2639–2642 CrossRefCASPubMed.
K. Bingol, F. Zhang, L. Bruschweiler-Li and R. Brüschweiler, Anal. Chem., 2013, 85, 6414–6420 CrossRefCASPubMed.
K. Bingol, F. Zhang, L. Bruschweiler-Li and R. Brüschweiler, J. Am. Chem. Soc., 2012, 134, 9006–9011 CrossRefCASPubMed.
Q. Zeng, J. Chen, C. Zhan, Y. Lin and Z. Chen, Anal. Chem., 2020, 92, 6893–6899 CrossRefCASPubMed.
J. Kikuchi, Y. Tsuboi, K. Komatsu, M. Gomi, E. Chikayama and Y. Date, Anal. Chem., 2016, 88, 659–665 CrossRefCASPubMed.
Y. Yan, M. T. Judge, T. Athersuch, Y. Xiang, Z. Liu, B. Jiménez and T. M. D. Ebbels, Anal. Chem., 2024, 96, 11707–11715 CrossRefCASPubMed.
R. T. Williamson, E. L. Chapin, A. W. Carr, J. R. Gilbert, P. R. Graupner, P. Lewer, P. McKamey, J. R. Carney and W. H. Gerwick, Org. Lett., 2000, 2, 289–292 CrossRefCASPubMed.
F. C. Schroeder, D. M. Gibson, A. C. L. Churchill, P. Sojikul, E. J. Wursthorn, S. B. Krasnoff and J. Clardy, Angew. Chem., Int. Ed., 2007, 46, 901–904 CrossRefCASPubMed.
R. R. Forseth, E. M. Fox, D. Chung, B. J. Howlett, N. P. Keller and F. C. Schroeder, J. Am. Chem. Soc., 2011, 133, 9678–9681 CrossRefCASPubMed.
R. J. Grams, W. L. Santos, I. R. Scorei, A. Abad-García, C. A. Rosenblum, A. Bita, H. Cerecetto, C. Viñas and M. A. Soriano-Ursúa, Chem. Rev., 2024, 124, 2441–2511 CrossRefCASPubMed.
H. Kim, J. Chin, H. Choi, K. Baek, T. G. Lee, S. E. Park, W. Wang, D. Hahn, I. Yang, J. Lee, B. Mun, M. Ekins, S. J. Nam and H. Kang, Org. Lett., 2013, 15, 5614 CrossRefCAS.
Y. Morishita, Y. Aoki, M. Ito, D. Hagiwara, K. Torimaru, D. Morita, T. Kuroda, H. Fukano, Y. Hoshino, M. Suzuki, T. Taniguchi, K. Mori and T. Asai, Org. Lett., 2020, 22, 5876–5879 CrossRefCASPubMed.
N. P. Paul, A. E. Galván, K. Yoshinaga-Sakurai, B. P. Rosen and M. Yoshinaga, BioMetals, 2023, 36, 283–301 CrossRefCASPubMed.
S. Hoshino, S. Ijichi, S. Asamizu and H. Onaka, J. Am. Chem. Soc., 2023, 145, 17863–17871 CrossRefCASPubMed.
A. E. Galván, N. P. Paul, J. Chen, K. Yoshinaga-Sakurai, S. M. Utturkar, B. P. Rosen and M. Yoshinaga, Microbiol. Spectr., 2021, 9, e0050221 CrossRefPubMed.
V. S. Nadar, J. Chen, D. S. Dheeman, A. E. Galván, K. Yoshinaga-Sakurai, P. Kandavelu, B. Sankaran, M. Kuramata, S. Ishikawa, B. P. Rosen and M. Yoshinaga, Commun. Biol., 2019, 2, 131 CrossRefPubMed.
W. Huang, J. Song, T. Sun, Y. He, X. Li, Z. Deng and F. Long, Nat. Commun., 2024, 15, 1659 CrossRefCASPubMed.
X. Zhang, F. Cheng, J. Guo, S. Zheng, X. Wang and S. Li, Nat. Synth., 2024, 3, 477–487 CrossRefCAS.
C. M. Kayrouz, K. A. Ireland, V. Y. Ying, K. M. Davis and M. R. Seyedsayamdost, Nat. Chem., 2024, 16, 1868–1875 CrossRefCASPubMed.
C. M. Kayrouz, J. Huang, N. Hauser and M. R. Seyedsayamdost, Nature, 2022, 610, 199–204 CrossRefCASPubMed.
F. Rossi, N. T. Duong, M. K. Pandey, M. R. Chierotti, R. Gobetto and Y. Nishiyama, Magn. Reson. Chem., 2019, 57, 294–303 CrossRefCASPubMed.
W. Gerrard, C. Yiu and C. P. Butts, Magn. Reson. Chem., 2022, 60, 1087–1092 CrossRefCASPubMed.
S. Cheatham, P. Gierth, W. Bermel and E. Kupče, J. Magn. Reson., 2014, 247, 38–41 CrossRefCASPubMed.
R. T. Williamson, A. V. Buevich and G. E. Martin, Tetrahedron Lett., 2014, 55, 3365–3366 CrossRefCAS.
L. Ciani and S. Ristori, Expet Opin. Drug Discov., 2012, 7, 1017–1027 CrossRefCASPubMed.
S. Hermanek, Chem. Rev., 1992, 92, 325–362 CrossRefCAS.
J. M. Macho, R. M. Blue, H. W. Lee and J. B. Macmillan, Org. Lett., 2022, 24, 3161–3166 CrossRefCASPubMed.
R. M. Blue, J. M. Macho, H. W. Lee and J. B. Macmillan, J. Nat. Prod., 2022, 85, 2682–2686 CrossRefCASPubMed.
H. Yu, H. Yang, E. Shi and W. Tang, Med. Drug Discovery, 2020, 8, 100063 CrossRefCASPubMed.
F. Bhinderwala, P. Evans, K. Jones, B. R. Laws, T. G. Smith, M. Morton and R. Powers, Anal. Chem., 2020, 92, 9536–9545 CrossRefCASPubMed.
M. A. DeSilva, N. Shanaiah, G. A. N. Gowda, K. Rosa-Pérez, B. A. Hanson and D. Raftery, Magn. Reson. Chem., 2009, 47, S74–S80 CrossRefCASPubMed.
K. S. Ju, J. Gao, J. R. Doroghazi, K. K. A. Wang, C. J. Thibodeaux, S. Li, E. Metzger, J. Fudala, J. Su, J. K. Zhang, J. Lee, J. P. Cioni, B. S. Evans, R. Hirota, D. P. Labeda, W. A. Van Der Donk and W. W. Metcalf, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, 12175–12180 CrossRefCASPubMed.
C. M. Kayrouz, Y. Zhang, T. M. Pham and K. S. Ju, ACS Chem. Biol., 2020, 15, 1921–1929 CrossRefCASPubMed.
Y. Zhang, L. Chen, J. A. Wilson, J. Cui, H. Roodhouse, C. Kayrouz, T. M. Pham and K. S. Ju, J. Am. Chem. Soc., 2022, 144, 9938–9948 CrossRefCASPubMed.
Y. Zhu, T. Shiraishi, J. Lin, K. Inaba, A. Ito, Y. Ogura, M. Nishiyama and T. Kuzuyama, J. Am. Chem. Soc., 2022, 144, 16715–16719 CrossRefCASPubMed.
L. Chu, X. Luo, T. Zhu, Y. Cao, L. Zhang, Z. Deng and J. Gao, Nat. Commun., 2022, 13, 1–16 Search PubMed.
L. F. Nothias, M. Nothias-Esposito, R. Da Silva, M. Wang, I. Protsyuk, Z. Zhang, A. Sarvepalli, P. Leyssen, D. Touboul, J. Costa, J. Paolini, T. Alexandrov, M. Litaudon and P. C. Dorrestein, J. Nat. Prod., 2018, 81, 758–767 CrossRefCASPubMed.
T. Nogawa, J. A. V. Lopez and H. Osada, in Cutting-Edge Organic Synthesis and Chemical Biology of Bioactive Molecules, 2019, pp. 1–9020 Search PubMed.
M. A. Cook, D. Pallant, L. Ejim, A. D. Sutherland, X. Wang, J. W. Johnson, S. McCusker, X. Chen, M. George, S. Chou, K. Koteva, W. Wang, C. Hobson, D. Hackenberger, N. Waglechner, O. Ejim, T. Campbell, R. Medina, L. T. MacNeil and G. D. Wright, J. Ind. Microbiol. Biotechnol., 2023, 50, kuad042 CrossRefCASPubMed.
J. Kjer, A. Debbab, A. H. Aly and P. Proksch, Nat. Protoc., 2010, 5, 479–490 CrossRefCASPubMed.
T. Grkovic, R. H. Pouwer, M. L. Vial, L. Gambini, A. Noël, J. N. A. Hooper, S. A. Wood, G. D. Mellick and R. J. Quinn, Angew. Chem., Int. Ed., 2014, 53, 6070–6074 CrossRefCASPubMed.
T. Grkovic, R. H. Pouwer, Q. Wang, G. P. Guymer, J. Holst and R. J. Quinn, J. Nat. Prod., 2015, 78, 1215–1220 CrossRefCASPubMed.
A. L. Harvey, R. Edrada-Ebel and R. J. Quinn, Nat. Rev. Drug Discovery, 2015, 14, 111–129 CrossRefCASPubMed.
C. A. Romero, T. Grkovic, J. Han, L. Zhang, J. R. J. French, D. I. Kurtböke and R. J. Quinn, RSC Adv., 2015, 5, 104524–104534 RSC.
D. Camp, R. A. Davis, M. Campitelli, J. Ebdon and R. J. Quinn, J. Nat. Prod., 2012, 75, 72–81 CrossRefCASPubMed.
U. Grienke, P. A. Foster, J. Zwirchmayr, A. Tahir, J. M. Rollinger and E. Mikros, Sci. Rep., 2019, 9, 11113 CrossRefPubMed.
J. L. Wolfender, G. Marti, A. Thomas and S. Bertrand, J. Chromatogr. A, 2015, 1382, 136–164 CrossRefCASPubMed.
V. Exarchou, M. Krucker, T. A. Van Beek, J. Vervoort, I. P. Gerothanassis and K. Albert, Magn. Reson. Chem., 2005, 43, 681–687 CrossRefCASPubMed.
W. S. Price, F. Elwinger, C. Vigouroux and P. Stilbs, Magn. Reson. Chem., 2002, 40, 391–395 CrossRefCAS.
S. H. Smallcombe, S. L. Patt and P. A. Keifer, J. Magn. Reson., Ser. A, 1995, 117, 295–303 CrossRefCAS.
R. M. Gathungu, R. Kautz, B. S. Kristal, S. S. Bird and P. Vouros, Mass Spectrom. Rev., 2020, 39, 35–54 CrossRefCASPubMed.
Y. Kobayashi and T. Shape, Cutting-Edge Organic Synthesis and Chemical Biology of Bioactive Molecules, 2019 Search PubMed.
Y. Lin, S. Schiavo, J. Orjala, P. Vouros and R. Kautz, Anal. Chem., 2008, 80, 8045–8054 CrossRefCASPubMed.
L. A. Trimble, M. W. Sumarah, B. A. Blackwell, M. D. Wrona and J. D. Miller, Tetrahedron Lett., 2012, 53, 956–958 CrossRefCAS.
A. W. Goering, R. A. McClure, J. R. Doroghazi, J. C. Albright, N. A. Haverland, Y. Zhang, K. S. Ju, R. J. Thomson, W. W. Metcalf and N. L. Kelleher, ACS Cent. Sci., 2016, 2, 99–108 CrossRefCASPubMed.
L. K. Caesar, F. A. Butun, M. T. Robey, N. J. Ayon, R. Gupta, D. Dainko, J. W. Bok, G. Nickles, R. J. Stankey, D. Johnson, D. Mead, K. B. Cank, C. E. Earp, H. A. Raja, N. H. Oberlies, N. P. Keller and N. L. Kelleher, Nat. Chem. Biol., 2023, 19, 846–854 CrossRefCASPubMed.
R. D. Kersten, Y. L. Yang, Y. Xu, P. Cimermancic, S. J. Nam, W. Fenical, M. A. Fischbach, B. S. Moore and P. C. Dorrestein, Nat. Chem. Biol., 2011, 7, 794–802 CrossRefCASPubMed.
R. D. Kersten, N. Ziemert, D. J. Gonzalez, B. M. Duggan, V. Nizet, P. C. Dorrestein and B. S. Moore, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, E4407–E4416 CrossRefCASPubMed.
B. Behsaz, E. Bode, A. Gurevich, Y. N. Shi, F. Grundmann, D. Acharya, A. M. Caraballo-Rodríguez, A. Bouslimani, M. Panitchpakdi, A. Linck, C. Guan, J. Oh, P. C. Dorrestein, H. B. Bode, P. A. Pevzner and H. Mohimani, Nat. Commun., 2021, 12, 3225 CrossRefCASPubMed.
L. Cao, A. Gurevich, K. L. Alexander, C. B. Naman, T. Leão, E. Glukhov, T. Luzzatto-Knaan, F. Vargas, R. Quinn, A. Bouslimani, L. F. Nothias, N. K. Singh, J. G. Sanders, R. A. S. Benitez, L. R. Thompson, M. N. Hamid, J. T. Morton, A. Mikheenko, A. Shlemov, A. Korobeynikov, I. Friedberg, R. Knight, K. Venkateswaran, W. H. Gerwick, L. Gerwick, P. C. Dorrestein, P. A. Pevzner and H. Mohimani, Cell Syst., 2019, 9, 600–608 CrossRefCASPubMed.
G. H. Eldjarn, A. Ramsay, J. J. J. Van Der Hooft, K. R. Duncan, S. Soldatou, J. Rousu, R. Daly, J. Wandy and S. Rogers, PLoS Comput. Biol., 2021, 17, 1–24 Search PubMed.
C. Xiang, S. Yao, R. Wang and L. Zhang, Beilstein J. Org. Chem., 2024, 20, 1476–1485 CrossRefCASPubMed.
F. Hemmerling and J. Piel, Nat. Rev. Drug Discovery, 2022, 21, 359–378 CrossRefCASPubMed.
H. Ren, C. Shi and H. Zhao, iScience, 2020, 23, 100795 CrossRefPubMed.
D. H. Nguyen, C. H. Nguyen and H. Mamitsuka, Briefings Bioinf., 2019, 20, 2028–2043 CrossRefCASPubMed.
K. Blin, H. U. Kim, M. H. Medema and T. Weber, Briefings Bioinf., 2018, 20, 1103–1113 CrossRefPubMed.
P. H. Willoughby, M. J. Jansma and T. R. Hoye, Nat. Protoc., 2014, 9, 643–660 CrossRefCASPubMed.
N. E. Avalon, M. A. Reis, C. C. Thornburg, R. T. Williamson, D. Petras, A. T. Aron, G. F. Neuhaus, M. Al-Hindy, J. Mitrevska, L. Ferreira, J. Morais, Y. El Abiead, E. Glukhov, K. L. Alexander, F. A. Vulpanovici, M. J. Bertin, S. Whitner, H. Choi, G. Spengler, K. Blinov, A. M. Almohammadi, L. A. Shaala, W. R. Kew, L. Paša-Tolić, D. T. A. Youssef, P. C. Dorrestein, V. Vasconcelos, L. Gerwick, K. L. McPhail and W. H. Gerwick, J. Am. Chem. Soc., 2024, 146, 18626–18638 CrossRefCASPubMed.
S. Khater, S. Anand and D. Mohanty, Synth. Syst. Biotechnol., 2016, 1, 80–88 CrossRefPubMed.
J. I. Tietz and D. A. Mitchell, Curr. Top. Med. Chem., 2015, 16, 1645–1694 CrossRefPubMed.
P. N. Leao, H. Nakamura, M. Costa, A. R. Pereira, R. Martins, V. Vasconcelos, W. H. Gerwick and E. P. Balskus, Angew. Chem., Int. Ed., 2015, 54, 11063–11067 CrossRefCASPubMed.
T. A. Scott and J. Piel, Nat. Rev. Chem., 2019, 3, 404–425 CrossRefPubMed.
E. J. N. Helfrich and J. Piel, Nat. Prod. Rep., 2016, 33, 231–316 RSC.
H. Mohimani and P. A. Pevzner, Nat. Prod. Rep., 2016, 33, 73–86 RSC.
L. Du, C. Sánchez and B. Shen, Metab. Eng., 2001, 3, 78–95 CrossRefCASPubMed.
M. G. Chevrette, F. Aicheler, O. Kohlbacher, C. R. Currie and M. H. Medema, Bioinformatics, 2017, 33, 3202–3210 CrossRefCASPubMed.
P. G. Arnison, M. J. Bibb, G. Bierbaum, A. A. Bowers, T. S. Bugni, G. Bulaj, J. A. Camarero, D. J. Campopiano, G. L. Challis, J. Clardy, P. D. Cotter, D. J. Craik, M. Dawson, E. Dittmann, S. Donadio, P. C. Dorrestein, K. D. Entian, M. A. Fischbach, J. S. Garavelli, U. Göransson, C. W. Gruber, D. H. Haft, T. K. Hemscheidt, C. Hertweck, C. Hill, A. R. Horswill, M. Jaspars, W. L. Kelly, J. P. Klinman, O. P. Kuipers, A. J. Link, W. Liu, M. A. Marahiel, D. A. Mitchell, G. N. Moll, B. S. Moore, R. Müller, S. K. Nair, I. F. Nes, G. E. Norris, B. M. Olivera, H. Onaka, M. L. Patchett, J. Piel, M. J. T. Reaney, S. Rebuffat, R. P. Ross, H. G. Sahl, E. W. Schmidt, M. E. Selsted, K. Severinov, B. Shen, K. Sivonen, L. Smith, T. Stein, R. D. Süssmuth, J. R. Tagg, G. L. Tang, A. W. Truman, J. C. Vederas, C. T. Walsh, J. D. Walton, S. C. Wenzel, J. M. Willey and W. A. Van Der Donk, Nat. Prod. Rep., 2013, 30, 108–160 RSC.
R. D. Kersten, N. Ziemert, D. J. Gonzalez, B. M. Duggan, V. Nizet, P. C. Dorrestein and B. S. Moore, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, E4407–E4416 CrossRefCASPubMed.
R. Reid, M. Piagentini, E. Rodriguez, G. Ashley, N. Viswanathan, J. Carney, D. V. Santi, C. Richard Hutchinson and R. McDaniel, Biochemistry, 2003, 42, 72–79 CrossRefCASPubMed.
U. Linne, S. Doekel and M. A. Marahiel, Biochemistry, 2001, 40, 15824–15834 CrossRefCASPubMed.
G. H. Hur, C. R. Vickery and M. D. Burkart, Nat. Prod. Rep., 2012, 29, 1074–1098 RSC.
M. C. Kim, H. Machado, K. H. Jang, L. Trzoss, P. R. Jensen and W. Fenical, J. Am. Chem. Soc., 2018, 140, 10775–10784 CrossRefCASPubMed.
K. Seshadri, A. N. D. Abad, K. K. Nagasawa, K. M. Yost, C. W. Johnson, M. J. Dror and Y. Tang, Chem. Rev., 2025, 125, 3814–3931 CrossRefCASPubMed.
T. F. Sun, Z. W. Ge, H. R. Xu, H. Zhang, S. S. Huang, M. G. Feng and S. H. Ying, J. Agric. Food Chem., 2024, 72, 18455–18464 CrossRefCASPubMed.
F. J. Ehinger, K. Scherlach, F. Trottmann, J. Fiedler, I. Richter and C. Hertweck, Angew. Chem., Int. Ed., 2025, 64, e202421760 CrossRefCASPubMed.
B. V. ija, B. Reddy, A. Milshteyn, Z. Charlop-Powers and S. F. Brady, Chem. Biol., 2014, 21, 1023–1033 CrossRefPubMed.
C. Wu, Z. Shang, C. Lemetre, M. A. Ternei and S. F. Brady, J. Am. Chem. Soc., 2019, 141, 3910–3919 CrossRefCASPubMed.
F. Y. Chang, M. A. Ternei, P. Y. Calle and S. F. Brady, J. Am. Chem. Soc., 2015, 137, 6044–6052 CrossRefCASPubMed.
E. J. N. Helfrich, R. Ueoka, A. Dolev, M. Rust, R. A. Meoded, A. Bhushan, G. Califano, R. Costa, M. Gugger, C. Steinbeck, P. Moreno and J. Piel, Nat. Chem. Biol., 2019, 15, 813–821 CrossRefCASPubMed.
J. Q. Hu, J. J. Wang, Y. L. Li, L. Zhuo, A. Zhang, H. Y. Sui, X. J. Li, T. Shen, Y. Yin, Z. H. Wu, W. Hu, Y. Z. Li and C. Wu, Org. Lett., 2021, 23, 2114–2119 CrossRefCASPubMed.
Y. H. Shin, J. H. Im, I. Kang, E. Kim, S. C. Jang, E. Cho, D. Shin, S. Hwang, Y. E. Du, T. H. Huynh, K. Ko, Y. J. Ko, S. J. Nam, T. Awakawa, J. Lee, S. Hong, I. Abe, B. S. Moore, W. Fenical, Y. J. Yoon, J. C. Cho, S. K. Lee, K. B. Oh and D. C. Oh, J. Am. Chem. Soc., 2023, 145, 1886–1896 CrossRefCASPubMed.
R. Alvarez and A. R. De Lera, Nat. Prod. Rep., 2021, 38, 1136–1220 RSC.
J. Park, Y. H. Shin, S. Hwang, J. Kim, D. H. Moon, I. Kang, Y. J. Ko, B. Chung, H. Nam, S. Kim, K. Moon, K. B. Oh, J. C. Cho, S. K. Lee and D. C. Oh, Angew. Chem., Int. Ed., 2024, 63, e202402465 CrossRefCASPubMed.
D. Shin, W. S. Byun, S. Kang, I. Kang, E. S. Bae, J. S. An, J. H. Im, J. Park, E. Kim, K. Ko, S. Hwang, H. Lee, Y. Kwon, Y. J. Ko, S. Hong, S. J. Nam, S. B. Kim, W. Fenical, Y. J. Yoon, J. C. Cho, S. K. Lee and D. C. Oh, J. Am. Chem. Soc., 2023, 145, 19676–19690 CrossRefCASPubMed.
Y. L. Du, H. Y. He, M. A. Higgins and K. S. Ryan, Nat. Chem. Biol., 2017, 13, 836–838 CrossRefCASPubMed.
K. D. Morgan, R. J. Andersen and K. S. Ryan, Nat. Prod. Rep., 2019, 36, 1628–1653 RSC.
K. D. Morgan, D. E. Williams, B. O. Patrick, M. Remigy, C. A. Banuelos, M. D. Sadar, K. S. Ryan and R. J. Andersen, Org. Lett., 2020, 22, 4053–4057 CrossRefCASPubMed.
K. D. Morgan, D. E. Williams, K. S. Ryan and R. J. Andersen, Tetrahedron Lett., 2022, 94, 153688 CrossRefCAS.
H. Huang, L. Yue, F. Deng, X. Wang, N. Wang, H. Chen and H. Li, J. Nat. Prod., 2023, 86, 2122–2130 CrossRefCASPubMed.
Y. Bando, Y. Hou, L. Seyfarth, J. Probst, S. Götze, M. Bogacz, U. A. Hellmich, P. Stallforth, M. Mittag and H. D. Arndt, Chem.–Eur. J., 2022, 28, 2–6 CrossRefPubMed.
H. Gross, V. O. Stockwell, M. D. Henkels, B. Nowak-Thompson, J. E. Loper and W. H. Gerwick, Chem. Biol., 2007, 14, 53–63 CrossRefCASPubMed.
Y. Yamada, T. Kuzuyama, M. Komatsu, K. Shin-ya, S. Omura, D. E. Cane and H. Ikeda, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, 857–862 CrossRefCASPubMed.
S. Chen, Q. Yang, X. Zhang, Z. Wang, H. M. Xu and L. Bin Dong, J. Nat. Prod., 2024, 87, 195–206 CrossRefCASPubMed.
D. W. Christianson, Chem. Rev., 2017, 117, 11570–11648 CrossRefCASPubMed.
D. A. Castillo, M. D. Kolesnikova and S. P. T. Matsuda, J. Am. Chem. Soc., 2013, 135, 5885–5894 CrossRefCASPubMed.
T. E. Smith, C. D. Pond, E. Pierce, Z. P. Harmer, J. Kwan, M. M. Zachariah, M. K. Harper, T. P. Wyche, T. K. Matainaho, T. S. Bugni, L. R. Barrows, C. M. Ireland and E. W. Schmidt, Nat. Chem. Biol., 2018, 14, 179–185 CrossRefCASPubMed.
Z. Pan and D. Raftery, Anal. Bioanal. Chem., 2007, 387, 525–527 CrossRefCASPubMed.
M. Wang, J. J. Carver, V. V. Phelan, L. M. Sanchez, N. Garg, Y. Peng, D. D. Nguyen, J. Watrous, C. A. Kapono, T. Luzzatto-Knaan, C. Porto, A. Bouslimani, A. V. Melnik, M. J. Meehan, W. T. Liu, M. Crüsemann, P. D. Boudreau, E. Esquenazi, M. Sandoval-Calderón, R. D. Kersten, L. A. Pace, R. A. Quinn, K. R. Duncan, C. C. Hsu, D. J. Floros, R. G. Gavilan, K. Kleigrewe, T. Northen, R. J. Dutton, D. Parrot, E. E. Carlson, B. Aigle, C. F. Michelsen, L. Jelsbak, C. Sohlenkamp, P. Pevzner, A. Edlund, J. McLean, J. Piel, B. T. Murphy, L. Gerwick, C. C. Liaw, Y. L. Yang, H. U. Humpf, M. Maansson, R. A. Keyzers, A. C. Sims, A. R. Johnson, A. M. Sidebottom, B. E. Sedio, A. Klitgaard, C. B. Larson, C. A. P. Boya, D. Torres-Mendoza, D. J. Gonzalez, D. B. Silva, L. M. Marques, D. P. Demarque, E. Pociute, E. C. O'Neill, E. Briand, E. J. N. Helfrich, E. A. Granatosky, E. Glukhov, F. Ryffel, H. Houson, H. Mohimani, J. J. Kharbush, Y. Zeng, J. A. Vorholt, K. L. Kurita, P. Charusanti, K. L. McPhail, K. F. Nielsen, L. Vuong, M. Elfeki, M. F. Traxler, N. Engene, N. Koyama, O. B. Vining, R. Baric, R. R. Silva, S. J. Mascuch, S. Tomasi, S. Jenkins, V. Macherla, T. Hoffman, V. Agarwal, P. G. Williams, J. Dai, R. Neupane, J. Gurr, A. M. C. Rodríguez, A. Lamsa, C. Zhang, K. Dorrestein, B. M. Duggan, J. Almaliti, P. M. Allard, P. Phapale, L. F. Nothias, T. Alexandrov, M. Litaudon, J. L. Wolfender, J. E. Kyle, T. O. Metz, T. Peryea, D. T. Nguyen, D. VanLeer, P. Shinn, A. Jadhav, R. Müller, K. M. Waters, W. Shi, X. Liu, L. Zhang, R. Knight, P. R. Jensen, B. Palsson, K. Pogliano, R. G. Linington, M. Gutiérrez, N. P. Lopes, W. H. Gerwick, B. S. Moore, P. C. Dorrestein and N. Bandeira, Nat. Biotechnol., 2016, 34, 828–837 CrossRefCASPubMed.
K. Bingol, L. Bruschweiler-Li, D. Li, B. Zhang, M. Xie and R. Brüschweiler, Bioanalysis, 2016, 8, 557–573 CrossRefCASPubMed.
K. Bingol, L. Bruschweiler-Li, C. Yu, A. Somogyi, F. Zhang and R. Brüschweiler, Anal. Chem., 2015, 87, 3864–3870 CrossRefCASPubMed.
C. Wang, B. Zhang, I. Timári, Á. Somogyi, D. W. Li, H. E. Adcox, J. S. Gunn, L. Bruschweiler-Li and R. Brüschweiler, Anal. Chem., 2019, 91, 15686–15693 CrossRefCASPubMed.
A. Leggett, C. Wang, D. W. Li, A. Somogyi, L. Bruschweiler-Li and R. Brüschweiler, Methods Enzymol., 2019, 615, 407–422 CAS.
K. Bingol and R. Brüschweiler, J. Proteome Res., 2015, 14, 2642–2648 CrossRefCASPubMed.
Y. Ge, C. Zhou, Y. Ma, Z. Wang, S. Wang, W. Wang and B. Wu, Anal. Chem., 2024, 96, 5399–5406 CrossRefCASPubMed.
A. A. Salim, K. Samarasekera, Z. G. Khalil and R. J. Capon, Org. Lett., 2020, 22, 4828–4832 CrossRefCASPubMed.
J. Gubbens, C. Wu, H. Zhu, D. V. Filippov, B. I. Florea, S. Rigali, H. S. Overkleeft and G. P. Van Wezel, ACS Chem. Biol., 2017, 12, 2756–2766 CrossRefCASPubMed.
C. Wu, M. H. Medema, R. M. Läkamp, L. Zhang, P. C. Dorrestein, Y. H. Choi and G. P. Van Wezel, ACS Chem. Biol., 2016, 11, 478–490 CrossRefCASPubMed.
P. A. Hoskisson and R. F. Seipke, mBio, 2020, 11, e02642 CrossRefCASPubMed.
M. W. Mullowney, K. R. Duncan, S. S. Elsayed, N. Garg, J. J. J. van der Hooft, N. I. Martin, D. Meijer, B. R. Terlouw, F. Biermann, K. Blin, J. Durairaj, M. Gorostiola González, E. J. N. Helfrich, F. Huber, S. Leopold-Messer, K. Rajan, T. de Rond, J. A. van Santen, M. Sorokina, M. J. Balunas, M. A. Beniddir, D. A. van Bergeijk, L. M. Carroll, C. M. Clark, D. A. Clevert, C. A. Dejong, C. Du, S. Ferrinho, F. Grisoni, A. Hofstetter, W. Jespers, O. V. Kalinina, S. A. Kautsar, H. Kim, T. F. Leao, J. Masschelein, E. R. Rees, R. Reher, D. Reker, P. Schwaller, M. Segler, M. A. Skinnider, A. S. Walker, E. L. Willighagen, B. Zdrazil, N. Ziemert, R. J. M. Goss, P. Guyomard, A. Volkamer, W. H. Gerwick, H. U. Kim, R. Müller, G. P. van Wezel, G. J. P. van Westen, A. K. H. Hirsch, R. G. Linington, S. L. Robinson and M. H. Medema, Nat. Rev. Drug Discovery, 2023, 22, 895–916 CrossRefCASPubMed.
V. J. Sahayasheela, M. B. Lankadasari, V. M. Dan, S. G. Dastager, G. N. Pandian and H. Sugiyama, Nat. Prod. Rep., 2022, 39, 2215–2230 RSC.
A. Gangwal and A. Lavecchia, J. Med. Chem., 2025, 68, 3948–3969 CrossRefCASPubMed.
K. Ito, X. Xu and J. Kikuchi, Anal. Chem., 2021, 93, 6901–6906 CrossRefCASPubMed.
J. Pollak, M. Mayonu, L. Jiang and B. Wang, Anal. Biochem., 2024, 695, 115654 CrossRefCASPubMed.
J. B. McAlpine, S. N. Chen, A. Kutateladze, J. B. Macmillan, G. Appendino, A. Barison, M. A. Beniddir, M. W. Biavatti, S. Bluml, A. Boufridi, M. S. Butler, R. J. Capon, Y. H. Choi, D. Coppage, P. Crews, M. T. Crimmins, M. Csete, P. Dewapriya, J. M. Egan, M. J. Garson, G. Genta-Jouve, W. H. Gerwick, H. Gross, M. K. Harper, P. Hermanto, J. M. Hook, L. Hunter, D. Jeannerat, N. Y. Ji, T. A. Johnson, D. G. I. Kingston, H. Koshino, H. W. Lee, G. Lewin, J. Li, R. G. Linington, M. Liu, K. L. McPhail, T. F. Molinski, B. S. Moore, J. W. Nam, R. P. Neupane, M. Niemitz, J. M. Nuzillard, N. H. Oberlies, F. M. M. Ocampos, G. Pan, R. J. Quinn, D. S. Reddy, J. H. Renault, J. Rivera-Chávez, W. Robien, C. M. Saunders, T. J. Schmidt, C. Seger, B. Shen, C. Steinbeck, H. Stuppner, S. Sturm, O. Taglialatela-Scafati, D. J. Tantillo, R. Verpoorte, B. G. Wang, C. M. Williams, P. G. Williams, J. Wist, J. M. Yue, C. Zhang, Z. Xu, C. Simmler, D. C. Lankin, J. Bisson and G. F. Pauli, Nat. Prod. Rep., 2019, 36, 35–107 RSC.
Footnote
† These authors contributed equally to this manuscript.