- Review
- Open access
- Published:
Solid-state NMR at natural isotopic abundance for bioenergy applications
Biotechnology for Biofuels and Bioproducts volume 18, Article number: 46 (2025)
Abstract
Lignocellulosic biomass offers a vast and renewable resource for biofuel production and carbon management solutions. The effective conversion of lignocellulosic biomass into economically competitive biofuels and bioproducts demands a comprehensive understanding of its complex structure and composition, often requiring a range of analytical tools to achieve meaningful insights. However, for the analysis of rigid solids, many traditional methods necessitate dissolution or chemical/physical modification of the sample, which limit our ability to capture an intact view of its structural components. This highlights the need for non-destructive approaches, such as solid-state nuclear magnetic resonance (ssNMR), which preserves the sample’s natural state while providing deep, molecular-level insights. While advanced multi-dimensional ssNMR on 13C-enriched materials has recently proven exceptionally valuable for elucidating the complex macrostructure of biomass, isotopic enrichment is expensive, laborious and is clearly infeasible at large scales. In this review, we explore the role of solid-state NMR methods at natural isotopic abundance as essential tools for the non-destructive, in-depth characterization of lignocellulosic biomass and bioenergy materials in their native and unaltered state. After a brief introduction to the basic principles of solid-state NMR, we first describe the acquisition and interpretation of routine 1D 13C ssNMR spectra of lignocellulose and other related biopolymers and products. We then delve into more advanced ssNMR approaches, including key spectral editing techniques, probing polymer dynamics, and various 2D methods applicable at natural abundance. Understanding of domain miscibility as observed from proton-based spin diffusion effects is a theme throughout. Our aim is to highlight key examples where ssNMR provides valuable insights into the composition, structure, dynamics, and morphology of rigid biomaterials relevant to the bioenergy economy, revealing both the native structures and fundamental transformations that occur across conversion and decomposition pathways. We hope that this review encourages a broader adoption of ssNMR methods in bioenergy research, where it can serve as a pivotal analytical tool for achieving sustainable biomass utilization and advancing a carbon-efficient bioeconomy.
Background
Lignocellulosic biomass, derived primarily from dedicated crops, agricultural waste, and forestry residue, exceeds 150 billion tons per year globally and presents enormous potential for the abundant production of biofuels and bioproducts from non-petroleum sources [1,2,3]. A wide array of strategies are being explored to overcome biomass recalcitrance and convert lignocellulosic biomass into fuels and products. For example, the conversion of biomass to traditional paper and packaging materials is a major global industry [4]. However, another major opportunity lies in second-generation biofuels, produced from non-food-based feedstocks like agricultural residues (e.g., corn stover, straw), forestry waste, and dedicated energy crops (e.g., switchgrass, poplar wood) [5, 6]. A current leading approach for second-generation biofuel production involves a three-stage process: (1) pretreatment to reduce recalcitrance and enhance enzymatic accessibility; (2) enzymatic hydrolysis to convert cellulose into fermentable sugars; and (3) fermentation to produce desired products like ethanol [7]. Common pretreatment methods include steam explosion and dilute acid or alkali treatments, with many alternative emerging strategies gaining attention including organosolv, ionic liquids, deep eutectic solvent pretreatment, mechanical refining, sonification, and microwave irradiation [8, 9].
A promising alternative strategy to the three-step biorefinery approach is consolidated bioprocessing, or CBP. By leveraging the natural biomass-degrading capabilities of thermophilic anaerobic microorganisms, like Clostridium thermocellum, CBP combines sugar production and fermentation into a single process potentially reducing costs and improving efficiency [10, 11]. Mechanical refining between sequential fermentations, referred to as cotreatment, has been proven effective at reducing certain biomass recalcitrance and enhancing total carbohydrate solubilization [12,13,14]. Also leveraging mechanical refining, the National Renewable Energy Laboratory (NREL) has pioneered a process called deacetylation and mechanical refining (DMR), which integrates low-severity chemical pretreatment and mechanical refining to boost sugar yields from lignocellulosic biomass while minimizing the production of fermentation inhibitors [15, 16]. This technology has gained attention because it can leverage equipment and processes similar to those used in the pulp and paper industry, reducing capital costs for biofuel production facilities.
Another leading pathway to biofuels production is thermochemical conversion strategies that utilize heat and chemical reactions to break down biomass into valuable products [17]. These methods include pyrolysis, which decomposes biomass at high temperatures in the absence of oxygen to produce bio-oil, biochar, and gases; gasification, where biomass is converted into syngas (a mixture of hydrogen, carbon monoxide, and carbon dioxide) through partial oxidation; and hydrothermal liquefaction, which uses high-pressure water to convert biomass into liquid biofuels as well as solid and gas products. These processes provide flexibility in producing a range of biofuels and co-products, including biochar, which has applications in materials, soil amendment and carbon sequestration [18].
In many of these approaches, as well as the pulp and paper industries, lignin is treated as a low-value waste byproduct. Lignin is the second most abundant polymer on earth after cellulose from terrestrial biomass and composes roughly 15–30% of the mass (feedstock dependent) and approximately 40% of the total energy value, but its complex and heterogeneous structure makes it challenging to valorize, and as a result it often ends up being underutilized. There are a number of promising strategies to handle lignin waste streams [19,20,21]. For one, lignin has high heating value and is often burned for process heating in biorefinery settings. While cost-effective, burning lignin waste streams fails to capitalize on lignin’s chemical potential to produce higher-value products. For example, lignin can be valorized directly in applications like biodegradable additives, lignin-based materials such as bioplastics and carbon fibers, and for long-term carbon storage solutions like incorporation into cement [22]. Alternatively, lignin-first biorefineries would first extract and depolymerize lignin into monomeric units to produce valuable aromatic chemicals before the cellulose is converted [23, 24]. Reductive catalytic fractionation (RCF) is an exciting example of such a lignin-first approach, whereby lignin is extracted and selectively depolymerized into low molecular weight aromatic units that can be used as fuel blends or building blocks for bioplastics, resins and other specialty chemicals [25, 26].
In all such explorations, including polysaccharide-based materials, lignocellulose conversion to biofuels, understanding biomass decomposition, thermochemical processing and waste stream valorization, precise and comprehensive analytical characterization of solid residues is essential. Various analytical tools offer critical insights into the composition, variability, and macromolecular structure of biomass, while also monitoring the molecular transformations and changes in residual solids throughout the deconstruction and conversion processes. This detailed characterization of residuals enhances understanding of process efficiency and product quality, illuminating the retention and transformation of key structural components within the biomass. Inevitably this could include a plethora of techniques including wet-chemistry methods (e.g., compositional analyses, sugar release assays), chromatography (HPLC, GC), spectroscopic techniques (FTIR, UV–Vis, pyrolysis-MBMS, solution-state NMR), microscopy (optical, fluorescence, SEM, TEM), thermal characterization techniques (TGA, DSC) and scattering methods (XRD, neutron scattering) [27,28,29]. The National Renewable Energy Laboratory (NREL) has developed a series of Laboratory Analytical Procedures (LAPs) to standardize and validate such methods for analyzing biomass composition, which can be found freely available at nrel.gov. However, most common analytical characterization techniques (e.g., benchtop, wet chemistry, and many instrumental methods) require some kind of chemical or physical treatment or sample dissolution, only analyze a portion of the material, and are broadly incapable of revealing molecular-level interactions and physical structures responsible for recalcitrance.
Overcoming some of these limitations, solid-state nuclear magnetic resonance (ssNMR) methods provide compositional, structural, morphological, and dynamical information of solid and semi-solid samples in their native, unaltered state without the need for chemical or physical alteration, thus preserving the integrity and complex structure of the neat material. Advanced multi-dimensional ssNMR methods have proven exceptionally valuable in the last decade towards elucidating the molecular structure and architecture of 13C-enriched plant cell walls, including grasses, hardwoods, and softwoods and solid products derived from these feedstocks [30,31,32,33,34,35,36,37,38,39,40,41]. While details differ among feedstock types, ssNMR results on 13C-enriched biomass have led to a general consensus macromolecular architecture as follows: a nanoheterogeneous arrangement of cellulose bundles are encased in a sheath of decorated hemicelluloses (e.g., acetylated xylan) such that the evenly spaced hemicellulose decorations protrude away from the cellulose surface and provide a favorable surface for interactions with polyaromatic lignin.
However, isotopic enrichment of biomass is practically limited due to the high cost of 13C labeled CO2 and obvious challenges in producing mature lignocellulosic feedstocks with 13C enrichment at scales relevant to the bioenergy economy. Therefore, the aim of this review is to bring into focus the utility and application of ssNMR methods at natural isotopic abundance for revealing a fundamental understanding of lignocellulose, its deconstruction and conversion into marketable goods, and for various carbon management solutions (Fig. 1). We will first briefly review the fundamentals of solid-state NMR to provide necessary context for experimental methods and interpretation of cited examples. We then introduce the polymers and generalized 13C spectral features for lignocellulosics and related bioenergy materials detectible by ssNMR, then review first routine then advanced ssNMR techniques applicable for characterizing biomass and biomass-derived products and waste streams at natural 13C isotopic abundance. Generalized topics will include routine 13C spectra by cross-polarization or direct-polarization methods, spectral editing techniques, revealing dynamics-based information, select two-dimensional methods, and future directions. The interpretation of proton spin diffusion effects in relation to domain sizes and heterogeneities will be a theme throughout. As we drive towards an energy-secure bioeconomy and seek a fundamental understanding of the restrictions to biological and chemical conversion of biomass to fuels, chemicals, and various downstream products, we hope that this fresh perspective, in addition to other related summarizing ssNMR works [42,43,44,45,46,47], will guide researchers in the bioenergy space to apply solid-state NMR in addition to more traditional characterization techniques expanding their analytical toolkit.
Overview of ssNMR at natural abundance for advancing bioenergy research. This figure underlines the crucial role of ssNMR in furnishing high-resolution information on molecular structure and chemical composition in biomass. ssNMR can give, through the resolution of important interactions among main lignocellulosic components, identification of structural motifs, cross-linkages, and dynamic behavior constituting the basis for understanding biomass recalcitrance. It is upon such insights that further optimization of biomass transformation processes may be founded and therefore ease the road toward the much-wanted development of sustainable energy technologies and renewable bio-based materials
Main text
Fundamentals of solid-state NMR
NMR is a sophisticated analytical technique that exploits the resonance of atomic nuclei when placed in an external magnetic field and exposed to electromagnetic radiation at specific frequencies. Most elements of the periodic table have at least one isotope that is NMR-active (e.g., spin quantum number is not 0), and are thus fundamentally detectable by NMR. In biomaterials research the most commonly used nuclei are 1H, 13C, and sometimes 15N, 31P and 2H. 1H is the most abundant and has the highest gyromagnetic ratio, making it the most common NMR nucleus for liquid-state NMR, but it is not commonly observed directly in the solid state due to strong and detrimental dipolar interactions (Fig. 2a). Atomic nuclei with a net charge generate a magnetic field due to their intrinsic spin angular momentum, which produces a magnetic moment (µ). This magnetic moment arises from the nucleus’ intrinsic spin, which can be visualized as the nucleus rotating around its own axis. When a nuclear magnetic moment is placed in an external static magnetic field, it results in a net magnetic moment (µ0) within the sample, referred to as net magnetization (M0). The magnitude of the detectible net magnetization is influenced by factors such as the number of detectable spins (N), the strength of the external magnetic field (B0), the gyromagnetic ratio (γ) and spin quantum number (I) of the nucleus, and the sample temperature (T) (Fig. 2b).
Technical aspects of ssNMR spectroscopy. a Commonly utilized NMR-active nuclei in biomolecular research, highlighting their relative sensitivities and natural abundances. 13C and 1H nuclei are predominantly employed in studies of lignocellulosic materials. b NMR interactions in solids: overview of key NMR interactions such as Zeeman, dipolar couplings, and chemical shielding interactions. These ranges highlight typical observations, though variations may occur based on molecular structure and local environments. c Chemical shift range for 13C nuclei, spanning from 0 to 220 ppm, utilized for structural elucidation. d Diagram illustrating the relationships among the magic angle (θr), the laboratory frame, the rotor frame, and the principal axis system, with the sample rotor oriented at the magic angle of 54.7°. e Representative solid-state NMR pulse sequences of initial polarization, including direct polarization (DP), cross-polarization (CP), and dynamic nuclear polarization (DNP). The asterisk (*) indicates RAMP-CP, which mitigates rf inhomogeneities and enhances polarization transfer efficiency. The ramp can be applied on either the 1H or X channel. f Modified CP sequences for side-band suppression and quantitative detection. g 1D 13C NMR spectra of glycine acquired under static conditions and at various magic angle spinning (MAS) rates. The asterisks (*) denote spinning sidebands, which decrease in intensity with increasing MAS frequency, leading to enhanced resolution of the isotropic chemical shifts. h 1D 13C NMR spectra of glycine recorded with and without heteronuclear dipolar decoupling. The spectrum without decoupling exhibits significant line broadening due to 1H–13C dipolar interactions, while the application of decoupling yields sharper 13C resonances, enhancing spectral resolution. All experiments in g and h were conducted on a 600-MHz spectrometer with a 4-mm HXY Phoenix probe, under 13.5 kHz MAS at 290 K, with swept-frequency two-pulse phase modulation (SWf-TPPM) heteronuclear dipolar decoupling sequence at the NREL
To manipulate the bulk magnetization vector and probe atomic environments, radiofrequency (rf) pulses are employed to apply on the samples. These pulses vary in field strength (B1) and frequency, typically near the Larmor frequency of the nuclei.
NMR spectroscopy provides profound insights into molecular and material structures by exploiting several key principles. In addition to the main factors of nuclei and strength of external magnetic field (Larmor frequency) and quantity (number of spins), NMR signals are also governed by the following key nuclear interactions: scalar couplings (J) between adjacent nuclei that give rise to multiplicities, chemical shielding (CS) interactions that depend on the local electronic shielding environment of the nucleus, space-based homonuclear and heteronuclear dipole–dipole interactions (D) between nuclei, and quadrupolar interactions (Q). The combined features result in deeply rich spectroscopic data reflecting the quantity, electronic environment (i.e., functional groups) and chemical bonding (Fig. 2b). Of particular importance is the sensitivity of NMR frequencies to the electron distribution around the nuclei. This chemical shielding interaction alters the nuclei's magnetic environment, resulting in slight frequency variations due to differences in the shielding constant (σ) among nuclei of the same isotope within a molecule. Frequencies are reported as chemical shifts (δ) in parts per million (ppm), with reference compounds like tetramethylsilane (TMS) and sodium trimethylsilyl propane sulfonate (DSS). Chemical shifts for 1H typically range from 0 to 20 ppm and for 13C from 0 to 220 ppm (Fig. 2c).
Challenges of NMR of rigid solids
NMR interactions are generally anisotropic, meaning the strength of these interactions depend on orientation of the magnetic moments with respect to the external magnetic field or with respect to other coupled nuclei. In the solution state where molecular tumbling is sufficiently fast, the dipole–dipole interactions average to zero as does the anisotropic component of the CS term, leaving only the isotropic component of the chemical shielding term. Intensities and multiplicities (J couplings) survive; thus we are left with a data-rich NMR signal that provides information on population (number of spins), chemical shielding environment, and any scalar couplings to adjacent nuclei. In the solid or semi-solid state however, where molecular tumbling is insufficient to average anisotropic nuclear interactions, the resulting NMR signal becomes severely broadened rendering it virtually impossible to interpret for most sample types. The characteristic values of one-bond dipolar coupling constants for 1H–1H and 1H–13C interactions are approximately 100 kHz and 22 kHz, respectively (Fig. 2b). It reflects the strong through-space dipolar interactions for nuclei at close proximity; the 1H–1H dipolar interaction is much stronger due to larger gyromagnetic ratio of protons. Further, in heterogeneous solids like biomaterials, the problem is compounded by the complex and varied molecular components, such as cellulose, hemicellulose, and lignin, which differ in crystallinity and chemical environment, further complicating the analysis and making it difficult to assign specific signals to individual molecular species without specialized techniques.
Key ssNMR techniques (MAS, CP, decoupling)
To overcome the limitations of NMR of rigid solids, three key techniques have been developed that are now standard practice: magic angle spinning (MAS), cross-polarization (CP), and high-power proton decoupling.
Magic-angle spinning (MAS) is a crucial technique developed in the 1950s to help overcome challenges associated with anisotropic interactions. By mechanically spinning the sample in a cylindrical rotor at the magic angle of 54.74° relative to the external magnetic field (Fig. 2d), MAS effectively averages out these anisotropic interactions leading to sharper, more interpretable spectral lines. This approach enhances the resolution and precision of ssNMR, providing clearer insights into the molecular and structural properties of solid samples.
Common pulse sequences for initial 13C polarization are shown in Fig. 2e. The simplest form of 13C ssNMR is direct polarization (DP), wherein the 13C nuclei are directly excited and observed. DP with MAS (DP-MAS) is often employed due to its straightforward approach, especially for non-rigid materials like solvent-swollen systems. However, it is crucial to acknowledge that 13C nuclei of highly rigid solids generally exhibit long longitudinal relaxation times (T1), which along with lower sensitivity (γC = ¼ that of γH) and low natural isotopic abundance (1.1%) severely limits the effectiveness of direct polarization. Circumventing these issues, the second major ssNMR breakthrough was the advent of cross-polarization (CP), introduced by Schaefer and Stejskal in the 1970s, where signal (magnetization) from an abundant spin like 1H is transferred to a dilute spin like 13C through the heteronuclear dipolar interaction [48, 49].
CP is widely utilized for two major reasons. First, CP is used to enhance the signal of low-γ nuclei, such as 13C and 15N, by exploiting the high gyromagnetic ratio and abundance of 1H spins in biological samples. Theoretical polarization enhancements are predicted to be γH/γX, which corresponds to approximately fourfold for 13C and tenfold for 15N. CP achieves polarization transfer by synchronizing radiofrequency irradiation on 1H and the observed 13C nucleus according to the Hartmann–Hahn (HH) condition (ɷ1,H = ɷ1,C ± nɷr), effectively locking the spins of the 1H and 13C nuclei in the transverse plane [50]. Second, with CP-based experiments one is limited by the relaxation delays of 1H instead of 13C, which are usually substantially faster and thus more signal averaging per unit time is possible. However, practical enhancements often fall short of theoretical values due to factors such as molecular dynamics and inhomogeneous polarization.
In CP, the HH matching profile can suffer significant mismatch issues when the MAS frequency approaches the sample’s dipolar couplings. This situation causes the profile to break into narrow bands separated by the spinning speed, complicating the achievement and maintenance of optimal matching conditions. Variable-amplitude cross-polarization (VACP) has been shown to effectively restore flat profiles at high spinning speeds [51]. A more advanced approach, the ramped-amplitude cross-polarization sequence (RAMP-CP) (Fig. 2e), further enhances this technique [52]. RAMP-CP enhances CP in ssNMR by varying rf field strength, broadening the HH matching condition, and improving 13C signal intensity. This mitigates rf inhomogeneities, ensures efficient polarization transfer, and enables more quantitative spectra. It also enhances resolution by reducing spinning sidebands, making it essential for complex material analysis.
The third key ssNMR building block is high-power proton decoupling, first introduced in the late 1950s and improved in subsequent decades [53, 54]. Here, strong radiofrequency fields are applied on the 1H channel to suppress dipolar interactions between abundant 1H nuclei and other nuclei, such as 13C. This decoupling technique eliminates or suppresses the influence of inter-nuclear couplings, which can broaden the spectra in solid samples, allowing for clearer detection of the less abundant nuclei. High-power decoupling is particularly critical in solid-state NMR, where strong dipolar couplings are prevalent, making it possible to observe sharp signals from nuclei like 13C even in complex materials like lignocellulosic biomass. The CP experiment can be modified with a series of 180° pulses to suppress sidebands, enhancing spectral clarity. Additionally, incorporating multiple CP blocks can improve quantitative accuracy by averaging polarization transfer across different molecular environments (Fig. 2f). More details are referred in the quantitative multi-CP and side-band suppression CP-TOSS section.
The combination of these techniques into cross-polarization magic angle spinning (CP-MAS) forms the foundation of modern solid-state NMR. To summarize and exemplify ssNMR on rigid solids like glycine, MAS reduces the broadening of signals by averaging out anisotropic interactions (Fig. 2g), while CP enhances the sensitivity of low-concentration nuclei like 13C and reduces experimental time thanks to the shorter 1H relaxation times, and proton decoupling sharpens the spectra by minimizing heteronuclear 1H–13C interactions (Fig. 2h). Together, these techniques allow for high-resolution analysis of solid materials like lignocellulosic biomass, revealing detailed molecular structures that are crucial for optimizing biofuel conversion processes. CP-MAS is commonly used as an initial step and integral component in various 1D, 2D, and 3D ssNMR pulse sequences, and further, the same concepts play a crucial role in transferring polarization from a more abundant sources, such as electrons, in indirect dynamic nuclear polarization (DNP) experiments (Fig. 2e).
CP-MAS, while widely used for signal enhancement, is not inherently quantitative due to its dependence on cross-polarization efficiency and dynamics. Quantitative solid-state 13C NMR spectroscopy is reliably and simply obtained using direct polarization DP-MAS, or with more advanced methods like multiple cross-polarization (MultiCP) [55]. DP-MAS is inherently quantitative, provided the recycle delay (d1) is set to 3–5 times the longest T1 relaxation time, ensuring complete spin–lattice relaxation, though it may require extended and sometimes untenably long acquisition times for carbons with long T1 values. In contrast, the MultiCP method offers an alternative for obtaining quantitative and high signal-to-noise ratio spectra with shorter acquisition times compared to direct polarization methods. The MultiCP pulse sequence works by repeated cross-polarization transfer from 1H to 13C, with repeat CP blocks separated by a relaxation delay, resulting in incremental magnetization buildup for all carbon types (Fig. 2f). These repeated CP steps cumulatively enhance the 13C signal, even for carbons with weak dipolar couplings to protons, such as non-protonated carbons. All of this put together allows MultiCP to reliably supply quantitative signal enhancement, thereby improving sensitivity and spectral resolution of all varieties of spin systems.
In 13C MAS NMR spectra, distinguishing between isotropic peaks and spinning sidebands can be challenging, especially when spinning sidebands overlap with isotropic peaks from other resonances. This overlap complicates spectral analysis. To mitigate this issue, an effective method is to acquire spectra at different spinning speeds. This allows the sidebands to shift in frequency, while the isotropic peaks remain stationary, making it easier to separate them. Alternatively, using a pulse sequence to suppress spinning sidebands can simplify the analysis by leaving only the isotropic peaks. A widely used sequence for this purpose is the TOSS (TOtal Suppression of Spinning Sidebands) sequence (Fig. 2f) [56]. It involves applying four precisely timed 180° pulses prior to FID acquisition, which randomizes the phases of the spinning sidebands while preserving the phase of the isotropic resonances.
These set of approaches, including CP-MAS, DP-MAS, MultiCP and TOSS comprise the set of routine 1D ssNMR applications. The core observable of a routine 13C MAS NMR experiment is to non-destructively reveal the carbon types of a solid sample regardless of structural configuration (amorphous, crystalline, semi-crystalline). While ssNMR suffers from plenty of disadvantages including low throughput, limited spectral resolution, inherently poor sensitivity of 13C at natural abundance and long acquisition times, one of the most significant advantages of solid-state NMR is its non-destructive nature, allowing lignocellulosic biomass and other bioenergy-relevant materials to be studied in its native, unaltered state. ssNMR makes it possible to explore the complex, multi-phase structure of these materials without the need for dissolving or chemically/physically modifying the sample. This is particularly important for lignocellulose and carbon management research where maintaining the integrity of the materials during analysis can provide more accurate insights into its natural structural properties and how they influence the efficiency of conversion processes. The following sections will first set the stage by conveniently introducing the major biopolymers detectable by ssNMR, then delve into first routine and then advanced ssNMR applications in the study of biomass and related bioenergy materials at natural isotopic abundance.
Biopolymers of lignocellulose and related bioenergy materials
Lignocellulosic biomass is a complex organic material that forms the structural framework of plant cell walls, primarily consisting mostly of cellulose, hemicellulose, and lignin. Cellulose, a polysaccharide composed of glucose units, provides strength and rigidity to the cell wall, while hemicellulose, a diverse group of polysaccharides such as xylans and mannans, binds closely to cellulose fibers, enhancing structural support and providing an interaction surface for lignin. Lignin, an aromatic hydrophobic polymer within the plant cell wall, adds rigidity and resistance to microbial degradation, making lignocellulose highly durable. Another important polysaccharide is pectin, which is found abundantly in the primary cell walls and middle lamella of plants. Acting as a glue, pectin plays a key role in cell adhesion and maintaining the structural integrity of plant tissues [57, 58].
Cellulose, one of the most renewable and abundant biopolymers, is a key component (30–50 wt%) of plant cell walls, enhancing their mechanical strength [59]. Discovered by Anselme Payen in 1838, it has the formula (C₆H₁₀O₅)ₙ, where “n” denotes the degree of polymerization. Composed of linear chains of β(1 → 4) linked d-glucose units (Fig. 3a), cellulose forms rigid, semi-crystalline microfibrils (CMF) of β-(1 → 4)-glucan chains, with diameters of 3–5 nm. Each cellulose chain features a d-glucose unit at one end, with a C4-OH as the reducing end and a C1-OH as the terminal group. In some forms, like bleached wood pulp, additional carbonyl and carboxyl groups may be present, modifying its properties [60].
Identified lignocellulose types in various biomasses. a Simplified schematic of cell wall polysaccharides, including cellulose, hemicellulose, and pectin. Only the glycosidic oxygens are shown while hydroxyl groups are omitted for clarity. Panel adapted with permission of ref [61]. b Schematic representation of the lignin structure, highlighting its three primary monomeric units: p-hydroxyphenyl (H, orange), guaiacyl (G, pink), and syringyl (S, brown). Ferulate (FA, light orange) and p-coumarate (pCA, cyan) are additional lignin-associated phenolic units predominantly found in grasses, typically ester-linked to hemicelluloses and lignin. The most common inter-unit linkages are emphasized in blue. Panel adapted with permission of ref [62]. c Schematic representation of the suberin structure, illustrating its aliphatic chains (blue) and lignin-like phenolic (yellow) compounds. Monomer units are shown with corresponding 13C chemical shifts labeled for each carbon site, with shifts derived primarily from references [34, 63,64,65,66]. d The relative abundance of cellulose, hemicellulose, and lignin, along with lignin compositional differences in each biomass, is shown
Hemicellulose is a heterogeneous group of polysaccharides in plant cell walls (approximately 15–35 wt%), comprising shorter, branched chains of sugars, such as xylose, mannose, and glucose (Fig. 3a). Hemicellulose often includes multiple sugar units and is categorized by its primary sugars, with key types including xyloglucans, xylans, mannans, glucomannans, and various other β-(1 → 3,1 → 4)-linked glucans [67]. Xylan, the most prevalent hemicellulose in terrestrial plants, is generally structured as a β-(1 → 4)-linked xylose backbone with varying substituents, exhibiting structural distinctions across plant types [67, 68]. In grasses and other non-grass monocots, xylan is primarily substituted with arabinose residues at the O-3 position of the xylosyl backbone, forming glucuronoarabinoxylan (GAX). These arabinose side chains can be further esterified by ferulic acid, which allows for cross-linking with lignin, contributing to the rigidity of grass cell walls. Additionally, grass xylan may include glucuronic acid (GlcA) side chains, though less frequently. In contrast, the xylan in woody dicots (e.g., hardwoods) is mainly glucuronoxylan (GX) is heavily acetylated, and is further decorated with α-(1 → 2)-linked glucuronic acid or 4-O-methylglucuronic acid residues attached at regular intervals to the xylan backbone [32]. This pattern is less complex than the branching seen in grass xylans. Both types can be acetylated, although the extent and pattern of acetylation vary, influencing how xylan interacts with other cell wall components. For example in hardwoods like poplar wood, xylan is acetylated every other xylose unit enabling the polymer to interact with the cellulose microfibril surface with acetate groups pointing generally away from cellulose, providing a favorable surface for lignin binding [31, 33, 41].
Pectin's structural complexity arises from its galacturonic acid backbone linked by α-(1 → 4) glycosidic bonds, with varying degrees of methyl esterification affecting solubility and gelling properties (Fig. 3a). It includes homogalacturonan (HG), rhamnogalacturonan-I (RG-I), and RG-II, which integrate into the cellulose–hemicellulose network to regulate wall porosity and cell expansion. Pectin also features neutral sugar side chains, such as rhamnose and arabinose, enhancing its three-dimensional structure [69].
Lignin exhibits remarkable structural complexity as a polymer composed of phenolic compounds, primarily formed from p-coumaryl, coniferyl, and sinapyl alcohols linked through various chemical bonds [70, 71]. The diversity in lignin's structure reflects its various polymers, functional groups, and linking methods. Three major types of phenylpropanoid building units can be identified: guaiacyl (G), syringyl (S), and p-hydroxyphenyl (H), corresponding to coniferyl, sinapyl, and coumaryl alcohols, respectively (Fig. 3b). Additionally, grasses contain notable amounts of ester-linked ferulate (FA) and p-coumarate (pCA) units, which contribute to cross-linking lignin with hemicelluloses [72]. Lignin also contains numerous functional groups, such as phenolic and alcoholic hydroxyls, methoxyls, carbonyls, and carboxyls, whose relative abundance varies by source, affecting its physical and chemical properties [73]. Unlike polysaccharides, lignin lacks a defined repeating unit, resulting in significant variability across plant species. This variability influences lignin's chemical reactivity and interactions with other biomolecules, while its degree of cross-linking enhances rigidity and resistance to degradation, providing structural support in cell walls and complicating biomass conversion in biofuel production. Additionally, lignin integrates into the cellulose–hemicellulose matrix, impacting the mechanical properties and accessibility of these components during degradation.
Both composition and structure vary significantly among differing feedstock types. By composition, hardwoods (e.g., poplar) generally contain about 40–50% cellulose, 25–35% hemicellulose-enriched xylose, and 20–30% lignin, which contains primarily guaiacyl and syringyl units, making them complex, dense, and durable in material structure. This composition and polymeric arrangements makes hardwoods structurally strong and stiff; thus, their main uses include furniture and cabinetry (Fig. 3d). The general composition of softwoods (e.g., pine) is 40–45% cellulose, 20–30% hemicellulose, mainly galactoglucomannan, and 25–35% lignin, with mostly guaiacyl units and no syringyl units. Importantly, these different compositions have a strong impact on the efficiency of enzymatic hydrolysis and fermentation processes, which all together may determine the sustainability of hardwood and softwood as feedstocks to produce biofuel and other bio-based materials. Lignocellulose in grasses, including wheat, rice, and switchgrass, is specifically constituted in ways that can enhance its use as a bioenergy feedstock. The composition generally consists of 30–50% cellulose, 20–30% hemicellulose rich in xylose and arabinose, and 10–25% lignin, mainly consisting of guaiacyl and syringyl units. The structural arrangement of biopolymers and their monomeric constituents, as well as other components, contributes toward the efficiency of hydrolysis with enzymes and fermentation. Thus, grasses represent one of the promising feedstocks in the realm of biofuels, even as they offer fast growth and adaptability.
Other biomass tissues in bark, roots, leaves and other plant organs consist of other biopolymers such as suberin and cutin. For example, cork, derived from the bark of the cork oak tree (Quercus suber), is a versatile biomaterial known for its lightweight, buoyant, and insulating properties. Composed mainly of 40% suberin, 22% lignin (predominantly guaiacyl and syringyl units), and 20% carbohydrates (including cellulose and hemicellulose), cork’s unique lignocellulosic composition provides water resistance and fire retardancy (Fig. 3d) [74]. Suberin is found in below-ground tissues like the root epidermis, endodermis, and periderm, and above-ground periderm from bark. This protective cell wall biopolymer is composed of phenolic and long-chain fatty acid monomers, with glycerol-based cross-linking. This is in contrast to other non-carbohydrate biopolymers, which include lignin and cutin, due to the presence of two unique domains: one poly(phenolic) and one poly(aliphatic), each with a distinct chemical composition (Fig. 3c). Both occur in the same cells, where they provide a barrier to prevent water loss, guard against pathogens, and participate in wound healing [75,76,77]. Cutin is a major constituent in plant leaves, flowers and fruit cuticles where it plays roles in water, solute and gas exchange as well as resistance to biotic and abiotic stressors [78]. Suberin and cutin are both lipopolymers that impart resilience to plant tissues and may play important roles in a bioeconomy that depends on sustainable plant traits and efficient use of all carbon-containing plant components [79].
Routine 1D 13C solid-state NMR of lignocellulose
Holistic overview of chemical shift regions and spectral features
Lignocellulosic compounds can be characterized and quantified using ssNMR while preserving their native state. The diverse chemical environments and functional groups present in various lignocellulosic materials result in several distinguishable chemical shift regions, allowing for effective differentiation and analysis of these complex structures, though with overlapping of signals existing at some levels. Here, we first provide a holistic overview of the 13C ssNMR fingerprint including the chemical shift ranges and structural features of typical biomass, then extend deeper into characterization applications aimed at tracking biomass conversion and deconstruction (Fig. 4).
Identified lignocellulose in biomasses using CP-MAS. a 1D 13C RAMP-CP spectrum of poplar, highlighting key chemical shift regions for lignocellulosic biomass: polysaccharides (50–100 ppm) and lignin (120–160 ppm). Illustration of NMR drive atomistic lignocellulose arrangement of Populus secondary cell wall. Panel adapted with permission of ref [41]. Illustration of C6 hydroxymethyl conformations in cellulose, with tg at 89 ppm (domain 1/crystalline/interior) and gg/gt at 84 ppm (domain 2/amorphous/surface). Assignments in overlapping regions are indicated with larger fonts for dominant species and smaller fonts for minor species. Panel adapted with permission of ref [80]. b 1D 13C RAMP-CP spectra of various biomass types, with suberin-specific chemical shifts (35–45 ppm) visible in the cork spectrum. The relative abundance of cellulose, hemicellulose, and lignin, along with lignin compositional differences in each biomass, is shown. All experiments were conducted on a 600-MHz spectrometer with a 4-mm HXY Phoenix probe, under 13.5-kHz MAS at 290 K, with SWf-TPPM heteronuclear dipolar decoupling sequence at the NREL
The 13C ssNMR spectra of biomass can be generally divided into four major spectral regions (aliphatic, carbohydrate, aromatic, carbonyl) with further sub-division if desired. The aliphatic region (~ 0–60 ppm) includes methyl and methylene carbons from hemicelluloses, protein backbone and sidechain, fatty acids and lipids, and methoxy carbons (Fig. 3d). The most characteristic ssNMR features for biomass include the acetate methyl at 22 ppm from acetylated hemicelluloses, fatty acid and lipid signals near 30 ppm, and lignin methoxy groups at 56 ppm. Suberin for example is rich in esterified fatty acids and thus reveal strong methylene features at 30 and 32 ppm. If pectin is abundant, for example in primary cell walls, a methyl ester group near 54 ppm is detectable, typically assigned to methyl-esterified galacturonic acid. Although minor in secondary cell walls, the rhamnose C6 methyl can often be seen near 17 ppm. The carbohydrate region exhibits characteristic spectral features within the range of 60 to 110 ppm. The anomeric carbon (C1) is typically deshielded relative to other non-anomeric carbohydrate signals and is observed as a cluster from approximately 98 to 105 ppm. This region is dominated by the cellulose C1 at 105 ppm, with an upfield shoulder from a heterogeneous mixture of hemicellulose environments. In poplar wood for example, the predominant hemicellulose (acetylated xylan) is multi-conformational with extended xylan (twofold, cellulose bound) exhibiting an anomeric carbon shift near 104 ppm while more disordered (threefold) C1 shifts coming in near 101–102 ppm. Other non-anomeric carbohydrate signals, for example cellulose C2,3,5 and xylan C2,3 overlap near 72–73 ppm.
Cellulose chemical shifts in NMR spectroscopy provide critical insights into its structural and dynamic properties, which are essential for understanding its role in biological and material systems. In solid-state 13C NMR, cellulose exhibits characteristic chemical shifts for its carbon atoms: C1 (~ 105 ppm), C4 (~ 89 ppm and ~ 84 ppm), C6 (~ 65 and ~ 62 ppm) and C2/3/5 (~ 75 and ~ 72 ppm). The split shifts of the C4 and C6 sites reflect the influence of hydrogen bonding, crystallinity, and molecular packing, particularly that of the hydroxymethyl C6 unit. Historically, the distinction between crystalline and amorphous regions was based on the 89/84 ppm chemical shift split, which was instrumental in developing the crystallinity index CrI (89/84 + 89) as a measure of cellulose order [42]. However, this simplistic interpretation is complicated for intact lignocellulose because the 84-ppm signal overlaps with contributions from matrix polymers, such as hemicelluloses and lignin inter-unit linkages, restricting the direct assessment of cellulose within intact biomass.
Recent advancements have introduced a more nuanced interpretation of cellulose chemical shifts. According to this refined interpretation, the 89 ppm signal reflects the more rigid packing structure of the hydroxymethyl groups in the tg conformation, which are predominantly found in the interior of the cellulose fibrils. In contrast, the 84-ppm signal corresponds to disordered hydroxymethyl groups in the gt/gg conformations, which are prevalent at the fibril surface. This understanding highlights that crystalline regions, which are more ordered, are predominantly located in the fibril core, while amorphous regions, which are more disordered, are more common at the surface. However to complicate matters further, the surface/interior assignment of 84 vs 89 ppm signals is contested by recent findings that a subfraction of surface cellulose can present at 89 ppm, leading to some authors assigning signals to domain 1 and domain 2 [30, 41]. A consensus is building that some surface cellulose may have its hydroxymethyl oriented towards the fibril interior, and it’s possible that in the densely packed secondary cell wall, surface cellulose units may hydrogen bond with matrix polysaccharides like bound xylan, stabilizing the tg conformation. To summarize, cellulose C4 at 89 ppm apparently arises when cellulose hydroxymethyl adopts a tg arrangement, which is more ordered or crystalline-like, and is over-represented but not exclusively within the cellulose microfibril core, while C4 at 84 ppm arises from a more disordered hydroxymethyl packing of gt/gg and appears more heavily represented on the fibril surface [80]. Therefore, assignments seen in the literature are mixed or interchanged: 89 ppm (crystalline, tg, interior, domain1) and 84 ppm (amorphous, gt/gg, surface, domain2).
Spectral overlap in the carbohydrate region of lignocellulose NMR spectra complicates cellulose crystallinity analysis, as signals from lignin inter-unit linkages overlap with cellulose carbons. For instance, the cellulose C1 resonance at ~ 105 ppm overlaps with hemicellulose C1 as well as syringyl lignin (S-lignin) C2,6. The carbohydrate signals in the 90–60 ppm range also overlap with lignin inter-unit linkages, for example the β-O−4 Cβ lignin linkage around 83 ppm, Cα near 73 ppm, and Cγ near 62 ppm. These complexities emphasize the need for advanced NMR techniques, such as spectral editing and 2D correlation or selective isotope labeling to differentiate cellulose from hemicellulose and lignin contributions.
The aromatic region spanning roughly 110–160 ppm is often dominated by lignin but can also include contributions from several other components depending on the biomass type or processing state. In addition to the characteristic lignin methoxyl group signals around 56 ppm, the aromatic region is information-rich regarding lignin’s composition and abundance. For example, syringyl units (S3/5) resonate distinctly at ~ 152 ppm while guaiacyl units (G3/4) appear around ~ 147–148 ppm [81]. Protonated ring carbons in guaiacyl units (G2/5/6) and typically resonate at ~ 110–120 ppm, while S2/6 overlap with carbohydrate anomeric carbons near 105 ppm. Para-hydroxyphenyl (H) lignin sometimes can be identified H3/5 at 116 ppm and H4 at 159 ppm. These distinct chemical shifts provide a structural fingerprint for assessing the relative contributions of S, G, and H lignin. In addition to lignin, hydroxycinnamates including ferulates and coumarates (common in grasses) and other polyphenolic compounds including tannins and flavonoids can also contribute to the aromatic region. Hydrolysable tannins and condensed tannins, such as proanthocyanidins, are abundant in certain biomass sources like bark, leaves, and some hardwoods, with signals from aromatic rings spanning ~ 110–160 ppm. Similarly, the aromatic domains of suberin contribute signals from phenolic components (~ 140–155 ppm), which are often linked to lignin-like structures. Minor contributions can also arise from the aromatic amino acids tyrosine and phenylalanine in proteins, which generate signals around ~ 115–135 ppm. These are typically present in small amounts but may become noticeable in protein-rich biomass and would be correlated with protein features in the aliphatic region. In biochars and other carbonized materials, the aromatic region reflects the presence of partially condensed or graphitic aromatic carbons. These signals, often broad and overlapping, are indicative of the degree of aromatic condensation and structural complexity. Overall, while lignin dominates in lignocellulose, the aromatic region in 13C NMR spectra serves as a complex landscape reflecting the diversity of aromatic and phenolic structures present in lignocellulosic and related biomass materials.
Lignin quantification and characterization
Lignin quantification and characterization are of utmost importance to understand the role of lignin in plant biomass, its interaction with the processing and utilization of lignocellulosic materials for bioenergy and bioproducts, and the valorization of biomass in a sustainable manner. Accurate lignin analysis would likely be very helpful on the development of strategy in biomass valorization such as enhancing enzyme hydrolysis efficiency, or optimizing fermentation processes.
There are several methodologies proposed for the quantification of lignin using 1D 13C CP spectra. Gao et al. developed a non-destructive quantification method of lignin using 13C CP-MAS spectra with an internal standard and spectral integration (Fig. 5a) [82]. A calibration curve of the quantification of lignin was obtained by using sodium-3-trimethylsilylpropionate (TMSP) as an internal standard in 13C CP-MAS. TMSP was chosen because it has a sharp peak at 0 ppm that is not overlapped by other peaks, and also because TMSP is soluble in distilled water, which facilitates its even dispersion. A 0.100 g/mL TMSP solution was prepared and added to each 250 mg biomass sample. In order to maintain homogeneous distribution, the samples were adjusted with distilled water or, when necessary, centrifuged for a short time. After freeze-drying, the resulting NMR spectra had given a linear calibration curve obtained by relating the lignin signal area to the TMSP peak. The method proved highly accurate and stable for the quantification of lignin, and its internal validation was approved.
Quantitative and characteristic insights of lignin. a Strategy of lignin quantification using internal standard. Chemical structure of internal standard is shown. Panel adapted with permission of ref [82]. b 1D 13C CP spectra of pure lignocellulose samples and cellulose-based sugarcane bagasse with calibration curve correlating the lignin contents with the area of the signal due to methoxyl groups in the 13C NMR spectra, obtained by deconvolution of the spectra. Panels adapted with permission of ref [83]. c Comparison of analytical methods for determining lignin unit ratios. Panels adapted with permission of ref [84]
Besides integration of the aromatic region, Cipriano et al. have used the well-resolved methoxyl group signal at 56 ppm for the quantification of lignin [83]. The 13C CP-MAS NMR spectra for pure lignin, cellulose, a cellulose–lignin mixture, and sugarcane bagasse (SB) provided resolutions representative of the distinct signals for lignin and cellulose and hemicellulose in SB (Fig. 5b). Overlapping carbohydrate signals complicate separation, but the methoxyl carbon signals in lignin give clear, quantifiable data for the lignin content. A calibration curve was established by correlating the methoxyl signal intensity with lignin mass in the cellulose–lignin mixtures. This allowed for accurate quantification of lignin in an array of biomass samples which were cross-validated with acid hydrolysis (AH) measurements. The majority of NMR-AH comparisons were consistent except in samples where carbohydrate hydrolysis was incomplete. Although the content of methoxyl groups does vary depending on the plant source, the methoxyl signal represents a very robust and reliable signal to quantify lignin in a wide range of biomass materials [85]. This method proved highly reproducible across several NMR instruments and thus finds wide applicability. Carbon signals from aromatics used simultaneously for the quantitative estimation of lignin were less effective due to poorer resolution and broader chemical shift ranges.
In addition to quantification, lignin composition can be partially interpreted by routine ssNMR. Previously, Manders et al. introduced non-destructive ssNMR to determine S/G ratios in hardwood by comparing spectra from hardwood, containing both S and G units, and softwood, which only contains G units [86]. This method, however, tends to underestimate the contribution of S-lignin due to normalization issues during spectral subtraction. Happs et al. addressed this by applying spectral deconvolution in the oxygenated non-protonated aromatic region (140–160 ppm) in 13C CP spectrum to improve the accuracy of S/G ratio calculations (Fig. 5c) [84]. They validated their method using poplar and found good agreement with results from thioacidolysis, py-MBMS, and HSQC. Although the ssNMR deconvolution approach is non-destructive, representative of the whole cell wall, and allows for the analysis of multiple phenotypes, it has drawbacks such as low throughput, sensitivity to incorrect peak fitting parameters (e.g., initial chemical shifts, line widths and Gaussian/Lorentzian ratios), and limited applicability to complex species like grasses. Nevertheless, it shows promise as an alternative to traditional methods for accurately quantifying S/G ratios, particularly in hardwood species.
Cellulose crystallinity
Cellulose crystallinity index (CrI) is an important factor that relate significantly to the physical and mechanical properties of cellulose-based materials, such as strength, stiffness, and water absorption. It is widely used to describe the relative portion of crystalline fractions in cellulose and quantify their modifications in plants and cellulosic materials following genetic engineering, and physicochemical or biological treatment. Solid-state NMR has been not only instrumental in discovery of two allomorphs of native cellulose, but also on measurement of cellulose crystallinity.
While the classical interpretations of cellulose crystallinity in ssNMR spectra (e.g., 89 vs. 84 ppm) is now better understood as more nuanced [80], studying the crystallinity of cellulose by ssNMR is nevertheless beneficial and informative. This is especially true for isolated cellulose preparations in which overlapping matrix polymers have been largely removed through processing efforts. Typically, crystallinity index is measured using isolated pure cellulose or extracted bacterial cellulose samples. Cellulose isolated from pulp, genetic engineered plants, and pretreated biomass are well-suited for CrI measurement using ssNMR. For example, Park et al. determined crystallinity index (CI) of Avicel PH-101 (a pure cellulose sample) used solid state using 13C CPMAS NMR by dividing the area of the crystalline peak (87–93 ppm) by the total area of the C4 peaks (80–93 ppm), a method well-suited for pure cellulose samples (Fig. 6a) [87]. The Ragauskas lab has documented a dataset of cellulose crystallinity index of 72 topline switchgrass natural variants using ssNMR spectroscopy [88]. We have also measured cellulose CrI of a total of an additional 54 various switchgrass lines, shown for the first time in Fig. 6b. For these studies pure cellulose samples were prepared for crystallinity measurement after removing the extractives, hemicellulose and lignin from the switchgrass, according to the published isolation procedure [89]. The cellulose crystallinity index across both sets was found in the range of 36–43% with the average of 39%, indicating only slight variation of cellulose crystallinity among the switchgrass natural population. It therefore appears that cellulose crystallinity estimates by ssNMR for isolated preparations only have a limited relationship to the recalcitrance difference in the biological conversion to sugars [90].
NMR characterization and assessment of cellulose crystallinity. a CPMAS spectra of Avicel PH 101 cellulose sample showing signals for crystalline and amorphous regions. Panel adapted with permission of ref [92]. b Crystallinity index by ssNMR for isolated cellulose preparations from topline switchgrass samples. NMR crystallinity measurements were carried out at Oak Ridge National Laboratory (ORNL) on a Bruker Avance-III 400 MHz spectrometer operating at a frequency of 100.59 MHz for 13C using a 4-mm double-resonance Bruker MAS probe. c CP-MAS of native poplar wood and flowthrough-pretreated poplar slurries using water. Panel adapted with permission of ref [91]
While ssNMR for cellulose CrI measurement have been largely applied on isolated cellulose samples, the cellulose isolation process is usually laborious and time-consuming. Therefore, ssNMR on characterization of plant biomass without isolation of cellulose that still contain other components has also been explored. For example Yan et al. studied the changes of whole biomass crystalline structure in poplar after flow-through pretreatment at various temperature based on signals in the C4 region of the 13C CP-MAS spectrum (Fig. 6c) [91]. This analysis revealed the gradual changes in C4 peaks corresponding to a degradation of amorphous components, enhancing our understanding of temperature effects on biomass crystalline structure disruption during flow-through. However, it should be noted that the crystallinity index measured using ssNMR for whole biomass is not cellulose CrI per se due to the overlapping signals contributed from other amorphous components like hemicellulose, lignin, or pectin. Spectral editing methods, described in later sections, can be useful for obtaining a cleaner look into the cellulose substructure for intact lignocellulose.
Tracking biomass deconstruction and conversion
Routine 13C ssNMR is not only invaluable for characterizing the native structure and composition of lignocellulose and its constituent components, but also serves as a powerful tool for tracking the transformations that occur during biomass deconstruction and conversion. The effective conversion of lignocellulose requires overcoming its natural recalcitrance, primarily due to the complexity of plant cell walls [93]. Various pretreatment and conversion methods including chemical, physical, biological, and thermal have been developed to enhance cellulose accessibility, lignin removal, and improving overall component fractionation. For example, fungal pretreatment offers strong degradability but may consume valuable carbohydrates and require longer processing times. Chemical pretreatment, such as using ionic liquids (ILs) and alkaline methods, provides high efficiency in lignin removal and cellulose accessibility but can generate inhibitors and require careful handling [94]. Meanwhile, bacteria and microbial consortia present adaptability and efficiency, though they face challenges in laboratory culturing and competition [95]. The following sections, loosely divided by processing method, describe key examples of utilizing routine 1D ssNMR (CP, MultiCP and DP-MAS) for monitoring spectral changes associated these processing and conversion steps, providing critical insights into how these treatments overcome biomass recalcitrance and enhance accessibility for bioenergy applications.
Chemical pretreatments
Chemical pretreatment strategies, including dilute acid, alkaline and ionic liquids, are among the more common ways to reduce recalcitrance and expose cellulose access for microbial conversion. As a clear example, Lima et al. studied the composition and effectiveness of five different pretreatments of various grassy feedstocks and bark varieties for the production of bio-renewables [96]. The biomasses contained 39–46% cellulose, 16–27% hemicellulose, and 21–25% lignin and were pretreated in hot water, acid, NaOH, and NaHSO₃ at different temperatures. Routine ssNMR analysis was applied to all feedstocks under various pretreatment conditions, demonstrating its power to non-destructively reveal structural and compositional changes. As a highlight, Fig. 7a shows how CP-MAS spectra tracked structural changes in sugarcane bagasse subjected to alkaline treatments at 50 °C and 130 °C. The spectra revealed that both temperatures effectively removed acetyl groups (22 ppm) and hemicelluloses (~ 70–80 ppm), but their impact on lignin removal differed significantly; at 50 °C, lignin removal was limited, as indicated by the persistence of aromatic signals (~ 110–160 ppm), while treatment at 130 °C showed enhanced delignification alongside increased cellulose crystallinity (~ 88 ppm). The ssNMR analysis also confirmed minimal compositional changes in samples pretreated with hot water or bisulfite, underscoring the limited efficiency of these methods. Among the pretreatments, it was shown that NaOH at 130 °C demonstrated the best performance in terms of enzymatic sugars release. These results suggest that NaOH will enhance saccharification more effectively and significantly reduce lignin under wide biomass type moderations of temperature. In a similar example applied to poplar wood, Sun et al. investigated the dilute acid pretreatment for the production of bioethanol and found that cellulose crystallinity and accessibility were increased, while enzymatic yield was higher in the delignified samples [97]. 13C CP-MAS showed that amorphous cellulose was preferably removed, promoting its recrystallization.
Analysis of chemical pretreatments using routine ssNMR. a Understanding novel biomass processing techniques for sugarcane bagasse. Panel adapted with permission of ref [96]. b Use of aprotic and protic ionic liquids for lignocellulosic biomass pretreatment. Panel adapted with permission of ref [98]. c HPAC pretreatment effectively targets and degrades amorphous cellulose. Panels adapted with permission of ref [99]. d Cellulose solvent pretreatment disrupts ordered hydrogen bonds in switchgrass fibers. Panels adapted with permission of ref [100]
In their research, Hossain et al. explored the effectiveness of imidazolium-based ILs containing acetate, formate, and chloride anions for pretreating lignocellulosic biomass [98]. The study examined both aprotic and protic varieties, utilizing quantitative 13C Multi-CP. Notably, the protic IL 1-ethylimidazolium chloride (EimCl) emerged as particularly effective, capable of dissolving whole wood (Pinus radiata) and significantly improving enzymatic hydrolysis of the biomass. In contrast, acetate- and formate-based protic ILs primarily served to extract lignin. Analytical techniques, including 13C Multi-CP and thermogravimetric analysis, revealed that EimCl selectively removes amorphous components from the biomass, thereby increasing cellulose crystallinity (Fig. 7b). This enhancement improves cellulase accessibility, leading to higher glucose yields. Meanwhile, the protic [Eim][OAc] primarily facilitates delignification without substantially altering cellulose characteristics [101]. These findings highlight the critical role of cellulose accessibility over mere lignin removal in effective biomass pretreatment. Additionally, NMR analysis indicates that acetate-based ILs significantly reduce the methoxy groups in lignin, suggesting structural changes, while chloride-based ILs have a milder impact on lignin structure. Overall, the type of anion and whether the IL is protic or aprotic are essential determinants of biomass pretreatment efficacy. Also applied to ionic liquid pretreatment with ssNMR, Husson et al. studied IL pretreatments, [Emim] + [CH3 COO] and [Emim] + [MeO(H)PO2] and found a significant decrease in the CrI of cellulose, while an increased glucose yield was obtained after [Emim] + [MeO(H)PO2] pretreatment due to increased accessibility of fibers [102].
Wi et al. evaluated the effectiveness of hydrogen peroxide-acetic acid (HPAC) pretreatment on pine wood biomass and were able to attain a lignin reduction of up to 98% with enhanced enzymatic digestibility (Fig. 7c) [99]. CP-MAS analysis indicated selective removal of amorphous components leading to higher glucose yields (93.4%) and better bioethanol production efficiency. Sathitsuksanoh et al. examined the cellulose solvent- and organic solvent-based lignocellulose fractionation (COSLIF) pretreatment, which combined phosphoric acid and organic solvents, significantly reducing CrI and enhancing cellulose accessibility through the degradation of amorphous regions (Fig. 7d) [100]. These studies demonstrate the effectiveness of NMR in monitoring lignocellulose structural and compositional changes induced by various pretreatments, which eventually improves cellulose deconstruction for biofuel production.
Physical impacts
Physical treatments, such as milling, extrusion, irradiation, sonication and other mechanical processes, are employed to reduce biomass particle size, and disrupt its structural integrity, and enhance polysaccharide accessibility and overall process efficiency. These physical treatment methods can be combined with other processing techniques for full and compounded effects, and their structural changes can be tracked using ssNMR. Ekwe et al. investigated the effect of ball milling on bamboo structure by using Multi-CP (Fig. 8a). Consequently, a treatment of dry biomass with 60 min of ball milling drastically reduced the crystalline cellulose content to about 10% in a crystalline state and increased the amorphous cellulose fraction. Lignin and hemicellulose showed less structural change (< 5%), indicating their stability during milling. Subsequent enzymatic hydrolysis partially removed hemicellulose and reduced amorphous cellulose while lignin remained intact. The shifting of the crystalline cellulose peak reflects a possible transformation of cellulose Iβ to Iα, while no formation of cellulose II was observed. NMR signal integration confirmed that the lignin content did not change after pretreatment and hydrolysis, indicating complete recovery of the sugars. These results demonstrate the effectiveness of mechanochemical pretreatment in disrupting cellulose crystallinity for the enhancement of enzymatic hydrolysis.
Analysis of physical pretreatments using routine ssNMR. a MultiCP spectra of bamboo showing the effects of ball-milling (top) and ball-milling with additional enzymatic hydrolysis. Panel adapted with permission of ref [104]. b Wheat straw pretreatment with ball milling and aqueous hydroxide solutions. Panels adapted with permission of ref [103]. c NMR spectra of cellulose showing chain-end signals and crystallite formation after treatment. The insert demonstrates the quantification of cellulose chain-end signals through spectral deconvolution. Panels adapted with permission of ref [105]. d Pretreatments induce distinct changes in the lignin aromatic region, reflecting structural modifications. Panel adapted with permission of ref [12]
Similarly, Qu et al. demonstrated that ultrafine ball milling of wheat straw prior to mild alkali treatment significantly enhances lignocellulose enzymatic deconstruction (Fig. 8b), with up to 98% glucose yield observed by enzymatic hydrolysis [103]. 13C CP-MAS revealed several key findings. First, ssNMR showed that NaOH treatment after extensive ball milling (ultrafine powder) was much more effective at removing lignin and hemicelluloses compared to NaOH treatment of un-milled material (< 1 mm and < 0.5 mm particle sizes). This can be seen by tracking acetate and lignin aromatic/methoxy signals before and after NaOH treatment at various particle sizes (Fig. 8b). CP-MAS spectra also reveal structural changes to cellulose upon aggressive ball milling. Substantial lineshape broadening of cellulose C1-C6 peaks was shown in ultrafine powder prior to alkaline treatment (ultrafine powder AT0), indicating disruption of inter- and intramolecular bonds during grinding, with signals at 62 and 84 ppm attributed to amorphous cellulose. Ultrafine powder also exhibited a unique saccharide oligomer signal near 97 ppm, which in hindsight could be assigned to cellulose end groups and therefore reduced degree of polymerization. Upon exposure to water (dilute NaOH) the cellulose recrystallizes, with ssNMR spectral features resembling cellulose II structure in addition to loss of lignin and hemicellulose.
Complimenting Qu’s work, a few years later Yuan et al. employed Multi-CP NMR to analyze Avicel PH-101 cellulose before and after ball milling followed by acid hydrolysis or ethanolysis, providing critical insights into cellulose structure (Fig. 8c) [105]. Signals at 97 and 92.7 ppm, in a 2:1 intensity ratio, were assigned to reducing chain ends, confirmed by their disappearance after ethanolysis, which forms ethyl ethers. Spectral shifts revealed cellulose-II crystallites with resonances at 107, 88, and 63 ppm, highlighting crystal modifications and reduced crystallinity. Despite these treatments, the chain-end signals remained unchanged, indicating their structural stability. In another example, Balch et al. demonstrated that one-pot microbial fermentation of switchgrass by Clostridium thermocellum, a process termed consolidated bioprocessing (CBP), can be enhanced by ball milling during fermentation. CBP augmented with milling is referred to as cotreatment, or C-CBP [12]. The authors report substantially higher total carbohydrate solubilization (TCS) when CBP is augmented with ball-milling (~ 88% TCS) compared to fermentations without cotreatment (~ 45% TCS). CP-MAS results (Fig. 8d) revealed minimal lignin modification caused by cotreatment, in contrast to hydrothermal pretreatment, which significantly alters lignin structure. A reduction in hydroxycinnamate moieties was also observed around ~ 116 ppm, suggesting their removal due to cotreatment. Interestingly, the authors reported a broadening of lignin signals, which presumably is attributed to residual metal contamination from milling with stainless steel hardware. Ball-milling with non-metal materials like zirconium oxide would eliminate this potential contamination. In addition to ball milling, 13C ssNMR methods have also been leveraged to characterize other physical pretreatment tactics including steam explosion [106,107,108], mechanical extrusion [109], and both microwave-assisted and ultrasound-assisted chemical pretreatments [110, 111].
Thermal conversion processes—biochar
CP-MAS and DP-MAS methods have also been utilized to analyze biochar chemistry based on differences in feedstock and conversion conditions, although there are challenges associated with concentration of metal content in biochars impacting NMR observability of these materials. Generally, residual biopolymer content (carbohydrates, lignin), functional groups and aromaticity are important features analyzed in biochars using CP-MAS techniques [43]. Methods such as 13C CP-MAS are being increasingly applied in the characterization of organic matter because of enhanced sensitivity and shorter measurement times. For example, Ben et al. performed a comprehensive 13C CP-MAS analysis on the temperature-dependent structural changes of biochar structures originating from different biomass sources (Fig. 9a) [112]. During low-temperature pyrolysis (400 °C), functional groups such as aromatic C–O, methoxyl-aromatic (~ 145 ppm), and aliphatic C–O bonds (~ 85 ppm) were preserved in the bio-chars. With increasing temperature, these functional groups undergo step-by-step decomposition. By 600 °C, bio-chars from cellulose, lignin, and tannin are transformed into condensed aromatic structures, reflecting the loss of side chains and oxygen-containing groups, accompanied by enhanced aromaticity and reduced structural complexity. Further, Brar et al. employed ssNMR among other techniques to investigate changes in biomass structure that determine the efficiency of saccharification after autohydrolytic and hydrothermal pretreatments [113]. Thus, native sugarcane bagasse, whose morphology is described by tightly packed fiber bundles with a coating of lignin, showed dramatic changes in morphology after pretreatment, that is, partial removal of the lignin layer due to delignification with lignin redeposition, thereby exposing fiber bundles (Fig. 9b). This morphological change increases the efficiency of enzymatic hydrolysis. Hydrothermal pretreatment was more effective compared with autohydrolytic treatment; this was probably due to the fact that hemicellulose is more effectively removed, and that the lignin structure is changed, offering the enzyme good accessibility to insoluble substrates. The most characteristic peaks in 13C CP-TOSS within the range between 50 and 120 ppm were assigned to cellulose although some contribution from hemicellulose and lignin was detected. It should be pointed out that peaks assignable to hemicellulose and lignin became weaker or even disappeared in the pretreated samples, which has been further supported through chemical compositional analysis. Furthermore, Chandel et al. used CP-TOSS to eliminate the overlap of spinning sidebands in the NMR, the former reported an increase of CI from 65.75 to 67.25 following enzymatic hydrolysis of SB [114].
Analysis of thermal pretreatments using routine ssNMR. a Temperature-dependent structural changes of biochar derived from different sources. Panels adapted with permission of ref [112]. b Hydrolysis enhancement of treated sugarcane bagasse showing structural changes for improved saccharification. Panel adapted with permission of ref [113]. c Natural and pyrolyzed rice husk: CP-MAS (9.4 T) highlights structures below 500 °C, while DP-MAS (2.0 T) emphasizes features above 500 °C. Stars indicate spinning sidebands. Panels adapted with permission of ref [115]
Freitas et al. also offered comprehensive insights from CP-MAS into the thermal transformation of lignocellulosic biomass, effectively distinguishing between crystalline (89.3 ppm) and amorphous (84.4 ppm) cellulose regions, hemicellulose acetates (21.8 and 174.3 ppm), and lignin-associated methoxyl (56.8 ppm) and aromatic carbons (115–150 ppm) [115].
However, CP-based methods have serious limitations in the analysis of biochars. The efficiency of 13C CP-MAS is poor for non-protonated carbons, mobile components, and fused-ring aromatic carbons in biochars, which results in considerable signal loss [43, 44]. Unlike CP-MAS, which depends on proton proximity, DP-MAS enables quantitative detection of all carbon types, making it particularly suited for analyzing carbonized materials with low hydrogen content. Shown in the same figure, Freitas pointed out the relevance of low-field 13C DP-MAS in the characterization of pyrolyzed lignocellulosic materials at high heat treatment temperatures (HTT) (Fig. 9c). High-field NMR, which gives excellent resolution to distinguish crystalline and amorphous cellulose, hemicellulose acetates, and lignin components, becomes inefficient at high HTTs due to the loss of hydrogen, where cross-polarization efficiency is weakened, hence leading to weaker signals, while aromatic carbons increase in chemical shift anisotropy (CSA), therefore leading to significant spinning sidebands, which obscure spectral resolution. Low-field NMR overcomes these issues by minimizing the CSA effects and providing much cleaner spectra dominated by an aromatic resonance at ~ 125 ppm, which indicates a highly ordered carbon matrix developed during pyrolysis. Direct polarization further helps to overcome the limitations of cross-polarization and thus allows for the robust characterization of highly carbonized materials. This makes lower-field DP-MAS NMR an important tool in understanding the structural transformation of lignocellulosic biomass under pyrolytic conditions. For a more comprehensive dive into ssNMR of biochars, please see the review by Baccile [43].
Biological treatments
Biological treatments, including fungal, microbial, enzymatic, and natural decay processes, play a pivotal role in biomass deconstruction by selectively targeting lignin, carbohydrates, and other structural components. Routine solid-state NMR techniques, particularly CP-MAS, have proven invaluable for characterizing the complex chemical and structural changes induced by these treatments. For example, CP-MAS was used by Sluiter and co-workers to help validate a new analytical protocol designed to quantify the abundance of cellulosic glucan in the presence of starch [116]. The protocol solves a challenging problem in biomass sugars compositional analysis for feedstocks in which both starch and cellulose are present, since both these glucan-derived polymers decompose into glucose monomers by two-stage acid hydrolysis and therefore the presence of starch can over-estimate the cellulosic glucan content. Accurately quantifying cellulosic versus starch-derived contributions to biofuel production is critical for ensuring compliance with regulatory frameworks and maximizing financial incentives, such as those provided under the Renewable Fuel Standard (RFS) and cellulosic-specific tax credits. Therefore, to accurately quantify cellulosic glucan within biomass when starch is also abundant, the authors proposed first chemically treating the sample with dilute NaOH to soften resistant starch granules, followed by complete enzymatic removal the starch with amylase while leaving cellulose unaffected. The CP-MAS results from Sluiter et al.’s study provided critical evidence validating the selective removal of starch from biomass while retaining cellulose (Fig. 10a). Specifically, the characteristic cellulose signals at 89 ppm and 105 ppm remained intact after dual chemical and enzymatic treatment, while the broad starch-associated resonance at 100–102 ppm was absent in the treated samples. The residual signal in this region was attributed to hemicellulose rather than starch, as confirmed by supplemental NMR methods.
Analysis of biological pretreatments using routine ssNMR. a Detecting the removal of starch under enzymatic and alkali treatment. Panels adapted with permission of ref [116]. b Variation of seven main organic carbon classes in 16 plant litters at different decomposition stages. Panels adapted with permission of ref [117]. c Kenaf raw samples after incubation with F. pinicola (brown rot) and G. lucidum (white rot). Panels adapted with permission of ref [85]. d Natural decomposition dynamics of hardwood leaf litter by Ganoderma australe. Panels adapted with permission of ref [118]. e Decay of Colorado Blue Spruce by Postia placenta. Panels adapted with permission of ref [119]
CP-MAS has also provided valuable insights into the natural microbial decay of biomass, for example from soil microbiota or fungi, revealing selective lignin degradation, carbohydrate removal, and structural transformations that govern decomposition rates and biomass digestibility. For example, Bonanomi et al. utilized 13C CP-MAS for the detailed chemical characterization of plant litter, demonstrating its superior predictive power for decomposition rates compared to conventional indices [117]. The CP-MAS spectra enabled the quantification of carbon functional groups, including alkyl carbons (lipids, waxes), O-alkyl carbons (carbohydrates), aromatic carbons (lignin-derived), and carbonyl carbons (carboxylic acids, esters). Spectral integrations are shown in the bar chart in Fig. 10b. A key finding was the strong correlation between the relative abundance of alkyl and O-alkyl carbons and litter decay rates, with alkyl-rich components being highly recalcitrant and O-alkyl-rich fractions more labile and easily degraded by microbial activity. This chemical profiling surpasses conventional metrics such as carbon-to-nitrogen content and lignin/N ratios by capturing the heterogeneity of carbon structures influencing decomposition [120, 121].
Lignin-degrading fungi that have been used in biomass deconstruction include Ganoderma lucidum and Fomitopsis pinicola. Fu et al. developed a method of quantitatively measuring the amount of lignin using 13C CP-MAS by producing a standard curve using integral values from 157 to 142 ppm (Fig. 10c) [85]. With this curve, these authors were able to measure the amount of lignin in untreated kenaf biomass at 32% [122]. This research further demonstrated that lignin degradation via enzyme can also be measured quantitatively. In particular, treatment using white-rot fungi G. lucidum was effective in reducing the lignin content in kenaf by more than 16% to 26%, while brown-rot F. pinicola did not greatly influence the change in kenaf lignin content.
Martinez et al. employed 13C CP-MAS to investigate the progressive lignin degradation in Eucryphia cordifolia wood by G. australe [118]. The spectra, shown in Fig. 10d, reveal pronounced changes in aromatic C signals with partial delignification at the medium decay stage and near-complete lignin removal at the advanced stage (Klason lignin ~ 6%), accompanied by increased cellulose content and in vitro digestibility. Notably, the carbohydrate signals remain consistent throughout the decay, indicating selective lignin degradation without significant alteration to the polysaccharide matrix, which likely enhances biomass digestibility. Lastly, Fig. 10e shows work by Davis et al., who employed 13C CP-MAS to investigate fungal degradation (brown rots) of spruce wood, revealing selective carbohydrate removal and complex lignin transformations [119]. Enhanced aromatic signals (110–160 ppm) and diminished carbohydrate peaks (105.7, 63, and 84 ppm) highlight the preferential loss of hemicellulose, while new carbonyl peaks (220–165 ppm) indicate oxidative changes. The study highlights the presence of lignin modifications, including reduced methoxy groups, increased phenolic hydroxyls, and demethylation, without significant aromatic ring cleavage.
Advanced 1D: spectral editing techniques
Considering that lignocellulosic-derived materials are complex and heterogeneous multi-polymeric systems, characterization by ssNMR methods is somewhat limited due to spectral overlap and poor resolution. Spectral editing techniques can provide useful simplifications of complex spectra. In the following section, we introduce and review the more common methods by which selection and editing of 13C ssNMR spectra can be obtained.
Dynamics-based filtering
Relaxation and dynamics filtering methods are particularly useful for distinguishing mobile-like or rigid-like carbon environments. Generally, these techniques take advantage of differences in laboratory frame longitudinal (T1) or transverse (T2), or rotating frame (T1ρ) relaxation of either the directly observed 13C nucleus or of the proton spins prior to CP. Molecular motion on either the nanosecond or microsecond timescales are relevant; spin–lattice relaxation in the laboratory frame (T1) are maximally sensitive to molecular motion on the scale of the Larmor frequency (MHz timescale), while relaxation in the rotating frame (T1ρ) is sensitive to motion near the spin-lock frequency, typically in the kHz regime [123].
Figure 11a shows the generalized T1/T2 curve as a function of the rotational correlation time. Generally, spin–lattice (T1) relaxation undergoes a T1 minimum in the intermediate dynamics regime when the correlation time is on the same order as the Larmor frequency (MHz, or nanosecond timescale). According to Ghosh and Dey, general “global” correlation times of cellulose are on the order of 1E-5 for amorphous or 1E-6 for crystalline environments [125]. Other studies on intact lignocellulosic materials suggest that the global dynamics of matrix polymers are of similar order as amorphous cellulose moieties [31, 34, 41]. At modern spectrometer frequencies of ~ 5–15 Tesla (200–600 MHz), these correlation times would place most if not all 13C environments within the lignified plant secondary wall on the ‘solid-side’ of the T1 curve. Thus, as the molecular motion increases, we typically observe a decrease in the T1 spin–lattice relaxation times (faster relaxation) owing to an increase in the molecular motion at or near the Lamour frequency. DP-MAS methods are capable of observing all carbon types regardless of molecular motion, but polymeric carbon environments with relatively increased molecular motion are therefore emphasized using a fast recycle delays, since slowly relaxing sites (e.g., rigid cellulose) become saturated and therefore under-represented in the resulting 13C spectra. With this concept in mind, contrasting CP-MAS and DP-MAS results with varying CP contact times and/or recycle delay times is one of the more simple but powerful ways to directly emphasize rigid vs mobile carbon environments. Fast-recycle delay DP-MAS spectra can be compared to CP-based methods to, respectively, compare more mobile-like and rigid-like environments within the material, as demonstrated on water-hydrated switchgrass in Fig. 11a.
Introducing dynamics-based ssNMR measurements at natural abundance. a Example of detecting molecular regimes based on dynamics: DP detects mobile molecules using a short recycle delay to capture signals from flexible regions, while T₁-filtered CP selectively enhances signals from rigid molecules, enabling differentiation between mobile and rigid molecular domains. The small panel illustrates the relationship between NMR relaxation times and molecular correlation times, highlighting how variations in relaxation behavior correspond to different dynamic regimes of the sample. Experiments were conducted on a 600-MHz spectrometer with a 4-mm HXY Phoenix probe, under 13.5-kHz MAS at 290 K, at NREL. b Representative ssNMR pulse sequences: 13C-detected 1H T1ρ and 13C-detected 1H T1 inversion recovery. Asterisk indicates that 13C-detected 1H T1ρ can be measured site-specifically. c Proton spin relaxation editing enhances resolution and improves baseline clarity. The graph indicates the enzymatic sugar yields versus XNMR for all never-dried entries with various water/organic-cosolvent mixtures. Panel adapted with permission of ref [124]
Also leveraging differences in 13C T₁ relaxation times, Sparrman et al. developed a revised 13C solid-state NMR method for determining cellulose crystallinity [126]. Since crystalline and amorphous cellulose differ substantially in molecular mobility—and consequently in their 13C T1 relaxation rates—the authors applied a T1 filter following initial CP polarization. Using either inversion-recovery or saturation-recovery techniques, they selectively enhanced the crystalline cellulose subspectra. By integrating these filtered spectra and applying a simple equation, this method provided reliable and quantitative estimates of cellulose crystallinity.
Proton spin relaxation editing (PSRE)
Another form of dynamics-based filtering is proton spin relaxation editing (PSRE) ssNMR, a class of spectral editing techniques designed to filter the 13C CP-MAS profile based on differing relaxation rates (T1, T1ρ, T2) of the associated protons [127]. The general idea is to select particular carbon types based on the relaxation properties of the proton bath prior to 1H–13C cross-polarization. The first implementation of PSRE was by Zumbulyadis et al. to remove background signal from the rotor using a T1(H) filter [127]. Here, a 1H 180 inversion pulse is first applied followed by a carefully timed recovery delay such that 1H signal associated the rotor is nulled while desired signal remains inverted. A 1H hard pulse is then applied to rotate desired magnetization into the x–y plane for cross-polarization and 13C detection (Fig. 11b, 13C-detected 1H T1 by inversion recovery). Newman then showed that this methodology could be used to isolate the 13C spectra of heterogeneous systems based on differing T1(H) relaxation times [128]. It is important to note here that strong dipolar coupling between protons leads to efficient spin-diffusion, manifesting as 1H relaxation time averaging for spatially proximal domains (e.g., 30–50 nm for 1H T1). Thus, the separation of the 13C spectra by 1H T1 relaxation requires large phase-separated domains. The requirement of large domain sizes has restricted application of T1(H) filtering on lignocellulosic materials, since these systems are more homogeneous on the nanoscale based on advanced 13C ssNMR results [40, 41]. However, Preston et al. showed that PSRE with T1(H) filtering prior to 13C CP-MAS was capable of distinguishing decomposed (faster 1H relaxation) and undecomposed (slower 1H relaxation) wood chips, suggesting PSRE T1(H) filtering may be useful for monitoring spatially heterogeneous decomposition [129].
Rotating frame (T1ρ) relaxation filtering has proven more useful. At common (5–15 kHz MAS) spinning speeds 1H T1ρ is also sensitive to spin-diffusion and therefore averaging of relaxation times, but the length scale of spin-diffusion over the millisecond timescale typical of T1ρ measurements is roughly 2–3 nm or less [130], much smaller than for T1(H). Thus, selection of 13C signals is achievable through T1ρ(H) relaxation filtering if domain sizes are on the order of 2 nm or larger. Averaging or partial averaging of 1H T1ρ between unique carbon environments is indicative of generally homogeneous polymer mixing [131]. If spin-diffusion effects are not desired, for example if site-specific T1ρ filtering is intended, higher spinning speeds can be utilized to quench spin-diffusion and thereby T1ρ averaging [132]. Spin diffusion can also be suppressed by collecting the experiment under the Lee–Goldburg condition.
After initial 1H polarization, T1ρ(H) filtering is accomplished by adding a sufficiently long (~ 2–20 ms) 1H spin-locking period prior to cross-polarization. 13C signal associated with longer 1H T1ρ is retained while environments with shorter T1ρ may be severely attenuated or lost. Rotating frame relaxation is maximally sensitive to molecular motion at the spin-lock frequency, which is typically on the order of 50–100 kHz. For example, water-plasticized hemicellulose and associated domains may experience increased cooperative polymeric motion relative to cellulose on the kHz timescales and therefore faster T1ρ(H) relaxation, while lower-amplitude kHz cooperative motion for crystalline cellulose is reflected in longer T1ρ(H).
For this reason, a common implementation of T1ρ(H) filtering is to obtain cleaner cellulose subspectra of intact lignocellulose. This relates to our prior comments on interpretation of 1D 13C CP-MAS spectra of intact cellulose; correct interpretation of cellulose morphology (e.g., crystalline vs amorphous, or interior vs surface, or tg vs gt/gg) by ssNMR for intact cellulose is hindered by spectral overlap of cellulose signals with matrix polymers, especially in the ~ 84 ppm region. In an early implementation applied to various lignocellulose sources, Newman noted T1ρ(H) filtering of 13C spectra was key in estimating the weight-averaged lateral dimensions of cellulose crystallites by quantifying the crystalline/interior to amorphous/surface ratio [133]. It is important to note that by using T1ρ(H) filtering it is possible to overcome spectral overlap issues and estimate cellulose crystallinity for intact lignocellulose. Most notably, Walker used T1ρ(H) filtering of CP-MAS spectra to reveal cellulose crystallinity details of raw and pretreated lignocellulose from three bioenergy feedstock candidates sorghum, switchgrass and poplar wood [124]. These ssNMR-derived metrics were capable of predicting enzymatic sugar yields achieved by hydrolysis of residual cellulose, regardless of biomass type or organic pretreatment solvent used; an increase in measured cellulose crystallinity was correlated with an increase in scarification yield (Fig. 11c). The key, the authors argue, is that an increase in cellulose crystallinity measured by T1ρ(H) filtered 13C CP-MAS spectra is due to a decrease in the interactions with hemicellulose/lignin matrix at the microfibril surface. This finding is intriguing and possibly counterintuitive since amorphous cellulose is known to hydrolyze more easily than the crystalline allomorph [134]. Other attempts to correlate NMR-derived cellulose crystallinity estimates from isolated cellulose with enzymatic degradation efficiency have been unconvincing [135,136,137]. The discrepancy may be attributed to structural changes to cellulose during the isolation process; water has been shown to promote recrystallization of ball-milled cellulose faster than hydrolysis [138], and also pore structure collapse may restrict enzymatic access. Walker’s results suggest that cellulose crystallinity estimates by T1ρ filtered 13C ssNMR on intact, never-dried material could be a suitable predictor of enzymatic hydrolysis efficiency.
Dipolar dephasing
Another common form of spectral editing is dipolar dephasing, which can differentiate between rigid protonated carbons and non-pronated carbons or protonated but mobile carbons. The dipolar dephasing (DD) experiment, first implemented by Opella and Frey [143], is designed to filter complex spectra based on differences in 1H–13C dipolar interaction. The strength of the 1H–13C dipolar coupling is foremost directly related to the inter-nuclear distance and therefore highly sensitive to if the carbon nucleus is protonated or non-protonated. Implementation of the DD method consists of initial 13C polarization by direct or CP methods followed by a short (40–100 μs) period during which no 1H decoupling is performed (Fig. 12a). This leads to rapid dephasing of 13C signal with strong dipolar coupling to protons (i.e., rigid protonated sites) while signal from non-protonated carbons survives. In addition, since the dipolar term can be averaged by high molecular motion the technique is sensitive to local dynamics of the 1H–13C bond vector; protonated carbon sites with sufficiently fast molecular motion may survive a dephasing period. Protonated methyl groups, which undergo fast three-site hopping at normal temperatures, are therefore also typically observed. As a result, dipolar-dephased 13C spectra filters out rigid non-methyl protonated carbons while retaining non-protonated and dynamic carbons. To view primarily the rigid protonated carbon types one can take a difference spectrum between a standard CP spectrum and the DD spectrum.
Spectral editing reveals distinct carbon environments. a Representative ssNMR pulse sequences: 1H–13C dipolar dephasing CP and CAS filtered CP. b Suppression of non-quaternary resonances in fungi-treated Bermuda grass, highlighting quaternary carbons. Panel adapted with permission of ref [139]. c Wood-char aromatics exhibit a more condensed arrangement compared to lignin, resulting in faster diphase decay in lignin than in wood-char. Panel adapted with permission of ref [140]. d Detecting non-protonated (quaternary) carbons in loblolly pine wood. Panel adapted with permission of ref [141]. e Lyophilized hydrolysate from sugarcane bagasse contains lignin, observed through characteristic lignin resonances. Panel adapted with permission of ref [142]
There are several examples where dipolar dephasing is utilized to understand lignocellulose and its deconstruction. One of the first implementations was by Gamble et al., in which CP-MAS and dipolar dephasing were applied on Bermuda grass treated with differing lignin-degrading white-rot fungi (Fig. 12b) [139]. It was shown that P. chrysosporium targets polysaccharides and to a lesser degree phenolic components, while C. subvermispora more substantially removed phenolic components over polysaccharides, and further suggested C. subvermispora shows a preference for guaiacyl over syringyl units. In another early example, dipolar dephasing was used to highlight non-protonated carbons in carbonaceous materials including lignin as well as other natural organic matter like charcoal and humic acids [140]. In this work it was also shown that long dipolar dephasing delays can help identify large fused-ring systems, since large fused aromatic rings are devoid of protons so long dephasing times are required to recouple the 1H–13C dipolar interaction (Fig. 12c). Dipolar dephasing combined with TOSS was used to better characterize native and milled pine, with dipolar-dephased spectra showing improved clarity over routine CP. In particular, the authors emphasized that dipolar-dephased spectra helped distinguish between ketone (C=O) and aldehyde (H–C=O) units downfield of 170 ppm since as protonated C=O aldehydes are suppressed while ketones are less affected (Fig. 12d) [141]. Results showed that milling increased the concentration of carbonyl groups, particularly ketones, by breaking lignin’s ether bonds and generating new oxidated structures. Rezende used dipolar dephasing to aid in spectral assignments of the lyophilized hydrolysate after alkaline pretreatment of sugarcane bagasse [142]. Along with routine CP-TOSS spectra, DD results suggested that it is primarily lignin and minor cellulose that is removed on 4% NaOH pretreatment (Fig. 12e).
Rezende et al. also employed the CSA filter on the CPMAS-TOSS spectrum of hydrolysates from NaOH-pretreated sugarcane bagasse to selectively eliminate signals from cellulose and non-aromatic lignin carbons, thereby confirming the removal of cellulose during pretreatment (Fig. 12e) [142]. The CSA filter in solid-state NMR exploits the angular dependence of the CSA tensor to selectively enhance signals from nuclei with pronounced anisotropic shielding, such as sp2 carbons or aromatic systems. By combining specific pulse sequences, such as dipolar decoupling and rotor-synchronized recoupling, the CSA filter isolates the anisotropic contribution while suppressing signals from nuclei in isotropic environments, such as sp3 carbons (Fig. 12a). This approach is invaluable for probing the molecular orientation, electronic environments, and the structural anisotropy of rigid and crystalline systems, providing a deeper understanding of local electronic distributions and bonding geometries. This approach highlighted that the remaining lignin peaks were associated with high chemical shift anisotropy, particularly from aromatic and methoxyl carbons, whereas cellulose-related signals were effectively filtered out, validating the spectral assignments. The application of the CSA filter provided useful insights into the structural changes and removal of both cellulose and lignin during the alkaline pretreatment of the bagasse samples.
Furthermore, corn stover residuals were characterized throughout dilute acid pretreatment (DAP) and enzymatic conversion process with advanced ssNMR methods, including dipolar dephasing and other spectral editing tools [144]. The authors come to some possible explanations for limitations of biological conversion of corn stover; chemical modification of polysaccharides, accumulation of cinnamic acids in residuals, close association of carbohydrates and lignin, and adsorption of proteinaceous or other nitrogen-containing material. This work serves as an excellent case study for applying advanced ssNMR techniques at natural abundance to better understand the lignocellulosic conversion process.
Exploring 2D and pseudo-2D ssNMR at natural abundance
In the previous sections we overviewed one-dimensional ssNMR methods, and in particular routine 1D DP-MAS, CP-MAS, and spectral editing and dynamics filtering methods designed to simply complex ssNMR spectra. Sans isotopic enrichment or access to DNP hardware, multi-dimensional solid-state NMR methods are more challenging, especially for heterogeneous polymer mixtures. 13C–13C correlation experiments such as Proton Driven Spin Diffusion (PDSD) and J-Refocused INADEQUATE techniques are powerful tools for 13C-enriched lignocellulose, but they are all but out of the question at natural abundance. Nevertheless, the characterization of biomass by solid-state NMR is not restricted to 1D techniques. Here, we refer to two-dimensional (2D) methods as both pseudo-2D, where the second indirect dimension is directly sampled time information, and true 2D, where frequency-encoded spectral information is obtained after Fourier transform of the indirect dimension.
Psuedo-2D relaxation: proton relaxation measurements and sensitivity to domain mixing
Relaxation times obtained by solid-state NMR are particularly useful for probing dynamical phenomena as well as providing a window into domain mixing on the nanoscale. Many of the previously mentioned 1D filtering methods become pseudo-2D if multiple time-points (tau delays) are collected such that the second dimension is directly sampled time information. As a reminder, spin–lattice (T1) relaxation is maximally sensitive to molecular tumbling on the order of the Larmor frequency (MHz regime), while rotating frame (T1ρ) relaxation rates are maximally sensitive to molecular motions on the order of the spin-locking frequency, typically on the order of 50 kHz [123]. 1H T1 rates for rigid organic solids are usually measured indirectly via 13C through the inversion recovery tactic by varying the recovery delay τ from short to long, and 1H T1ρ can be measured by varying the spin-locking pulse length (Fig. 11b). Fitting the signal intensity versus time to a known exponential function yields the desired information. Importantly, 1H relaxation properties in the solid state are sensitive to proton spin diffusion resulting in an averaging of 1H relaxation times, and thus only domain-averaged information is usually accessible. Since 1H T1 times are typically on the order of seconds, and proton spin diffusion rate constants are on the order of 0.8 nm2/ms for rigid polymer systems [145], 1H T1 relaxation rates are expected to be averaged for homogeneous systems with domain sizes much smaller than ~ 30–50 nm. This can be both a feature and a bug. While site-specific 1H relaxation rates are not often obtainable, domain sizes and polymeric interactions can be loosely investigated through an understanding of 1H relaxation time averaging. For example, Ahvazi measured the 1H T1 for isolated lignin and bleached cellulose from cuoxam as a function of pH. The results showed very different T1(H) vs pH curves for cellulose and for lignin when isolated, but similar curves for the same lignin and cellulose signals within native wood, suggesting relaxation averaging due to polymer proximity [146]. Similarly, by comparing 1H T1 relaxation rates of native and denatured Japanese cypress wood in both the humid and heat-dried states, Nishida demonstrated that removal of lignin (and to a lesser degree hemicellulose) results in increased cellulose T1(H) times [147]. This observation is explained if lignin acts as the primary 1H relaxation sink for cellulose protons thanks to proton spin diffusion between the polymers. T1(H) analysis performed on native and steam-treated bamboo also showed a decrease in 1H T1 values with increasing temperature and treatment times, consistent with an increase in smaller and more dynamic pyrolysis products [108]. Proton T1 relaxation measurements have also demonstrated an increase on molecular mobilities of sulfonated and methylated softwood pulps when the sodium salt was used but the opposite for when the calcium cation was used [148]. As a cautionary note, it is clear that water hydration has a marked effect on 1H relaxation times [149], therefore moisture state should be reported if directly interpreting T1(H) for characterizing lignocellulose and derived materials.
Recall that spin–lattice relaxation times in the rotating frame (1H T1ρ) are in principle sensitive to molecular motions on the kilohertz timescales, with faster T1ρ generally associated with increased dynamics for lignocellulose. To highlight this phenomenon, 1H T1ρ analysis of relatively dry and hydrated onion primary cell walls revealed a substantial decrease in T1ρ for pectic sites but a minor decrease in T1ρ for cellulose sites upon hydration [150]. These results indicate water penetrates and plasticizes pectin to a much greater degree than cellulose. Also, T1ρ measurements help distinguish between Norway spruce (Picea abies) wood grown in three different geological regions. Samples from Finland showed more uniform and lower T1ρ values compared to those from Poland followed by Italy, and the authors suggested increased polymer mobility in the same order. Further, principal component analysis (PCA) was able to distinguish between the plot sights based on T1ρ rates [151].
Polymer mixing and spin-diffusion also play a role in rotating frame relaxation time averaging but on a much shorter length scale in the range of ~ 2 nm [130, 145]. Considering that polymers within lignocellulose are known to have sub-nanometer spatial interactions [30, 34, 40, 41], one must interpret T1ρ results from intact lignocellulose cautiously and expect some degree of averaging. For example, Newman et al. [152] showed similar T1p for crystalline and amorphous cellulose, and of hemicelluloses with lignin, suggesting intimate close-range interactions of the carbon types. T1p averaging was captured more indirectly by Leroy et al. [153] who studied the effects of hot water pretreatment (HWP) on maize stems and showed that both enzymatic saccharification efficiency and 1H T1ρ relaxation times were positively correlated with increased pretreatment time. Since for lignocellulose 1H T1ρ generally will increase with rigidity, at first it may seem counterintuitive that an increase in 1H T1ρ would be associated with increased enzymatic saccharification. But as matrix polymers are increasingly removed with pretreatment, 1H T1ρ averaging between more-rigid cellulose and less-rigid matrix polymers is reduced, thus increasing the observed 1H T1ρ of cellulose. Therefore, for intact lignocellulose an increase in 1H T1ρ may serve as a proxy for increased cellulose accessibility.
Relaxation: site-specific dynamics of lignocellulose
To obtain site-specific dynamics on the kilohertz timescales via 1H T1ρ, spin-diffusion effects must be quenched or absent. This is most commonly achieved using the Lee–Goldburg condition; by applying an LG spin-lock field instead of a standard spinlock to the 1H channel before the CP step. LG-CP is then used to transfer the 1H polarization to 13C in a site-specific manner. This approach ensures selective polarization transfer, so that 1H magnetization is efficiently relayed to directly bonded 13C nuclei, minimizing contributions from distant or weakly coupled spins. While we have not identified any examples of 1H T1ρ measurements applied to lignocellulose using the LG condition at natural isotopic abundance, its utility to 13C-enriched materials has proven enlightening [154, 155]. For example, in one of the earlier instances of advanced ssNMR of 13C-enriched plant materials, Dick-Perez and co-workers demonstrated that site-specific 1H T1ρ relaxation measurements reveal distinct dynamics of pectic polysaccharides in Arabidopsis thaliana primary cell walls [155]. Among the many findings from advanced ssNMR techniques, dynamics measurements highlighted the high mobility of pectins relative to cellulose and hemicellulose, and further demonstrated how depectination reduces molecular flexibility and increases polysaccharide packing density. Such techniques could be applicable at natural isotopic abundance.
Site-specific dynamics in the nanosecond/MHz timescales are also accessible through 13C since at natural abundance the opportunity for 13C–13C spin-diffusion and averaging is absent. Due to typically longer 13C spin–lattice relaxation rates compared to protons, measurement of 13C T1 by direct methods like saturation recovery or inversion recovery are often untenable. 13C spin–lattice relaxation times are therefore measured most commonly using the method developed by Torchia (Fig. 13a) [156], which uses 1H–13C CP for initial polarization along with phase cycling to achieve 13C signal decay based on 13C T1, but the repetition rate (time between scans) is dictated by the much faster proton relaxation.
Pseudo-2D and true 2D ssNMR techniques for biomass analysis. a Representative ssNMR dynamic pulse sequences: Torchia 13C T1 relaxation and the inset panel illustrates the correlation between the decay of NMR signal intensity over time and the corresponding relaxation time constant (T1 or T2). b Spin–lattice relaxation times for major cellulose carbon sites in the secondary cell walls of OsCAldOMT1- and OsCAD2-knockout rice mutant lines. Panel adapted with permission of ref [159]. c Representative ssNMR dynamic pulse sequence: 1H T2 filtered CP. d T2 filter followed by 1H spin diffusion then CP observation to understand water proximity. Experiments were conducted on a 600-MHz spectrometer with a 4-mm HXY Phoenix probe, under 13.5-kHz MAS at 290 K, at NREL. e Representative ssNMR true 2D pulse sequences: DIPSHIFT, WISE and 1H–13C HETCOR. Some sequences are collected with or without a spin diffusion block (indicated by a tSD mixing period) to investigate domain sizes and spatial arrangements
Here, we describe some key examples wherein 13C T1 rates are utilized to understand lignocellulose and its deconstruction at natural abundance. Ghosh used site-specific 13C T1 measurements on untreated cellulose from rice straw and sugarcane bagasse, and on their residuals post-enzymatic digestion [157]. The results suggested an overall decrease in 13C T1 times for all carbon types post digestion, pointing to an increase in disorder and possibly reduction in polymeric size due to enzymatic activity. Since the ratios of crystalline-like and amorphous-like cellulose appear unaltered, the authors argue that tracking 13C relaxation throughout lignocellulose deconstruction may be a more effective way of tracking changes in the cellulose structure than routine 1D 13C ssNMR. Martin and co-workers studied the chemical and ultrastructural effects of lignin-modified rice cell walls [158, 159]. For rice deficient in cinnamyl alcohol dehydrogenase (CAD2), site-specific 13C T1 relaxation times suggested a complex alteration of cellulose mobility and dynamics for the cad2 mutant compared to wildtype (Fig. 13b). Specifically, biexponential T1 analysis was applied to extract faster- and slower-relaxation components, and overall longer T1 relaxation was observed for mutant lines but also an increase in the relative fraction of faster-relaxing component was increased [158]. The authors propose that a chemical alteration of lignin via CAD deficiency has downstream mobility effects on cellulose due to an altered lignin–hemi-cellulose interplay. In a follow-up study, the authors added additional lignin-modified rice lines including 5-hydroxyconiferaldehyde O-methyltransferase (CAldOMT) and stacked CAldOMT1/CAD2 deficient lines, and showed via 13C T1 measurements that CAldOMT deficient and stacked CAldOMT/CAD deficient lines had an increase in cellulose mobility (reduced 13C T1) compared to wildtype [159]. Both CAldOMT1 and CAD2 lines demonstrated substantially enhanced enzymatic saccharification efficiencies relative to wildtype, although CAldOMT1 lines outperformed CAD2, suggesting a possible relationship between cellulose mobility and scarification efficiency. Other examples of applying 13C T1 analysis for understanding native and altered lignocellulose at natural abundance include probing the dynamic effects of dilute sulfuric acid pretreatment on poplar wood [160, 161], native and denatured bamboo [162], understanding water accessibility to cellulose [163], understanding of cellulose crystallinity [126], and using 13C T1 as a source of variability for multivariate analysis [164], to name a few.
Leveraging T2 relaxation for water proximity and domain sizes
The 1H T2 relaxation times for water and biomass vary significantly, with biomass protons showing much shorter 1H T2 times due to its more rigid and heterogeneous environment. By exploiting this difference, it is possible to selectively suppress the proton bath from biomass, leaving the initial magnetization primarily in water. If a 1H spin-diffusion period is included after 1H T2 filter, one can transfer 1H magnetization from water molecules to neighboring biomass components, then observe 13C signal after CP (Fig. 13c). Short tSD periods will highlight the carbon types of polymeric units immediately proximal to water. This approach can be used to probe the “hydration level” by mapping how closely water is associated with different biomass domains based on proximity-dependent polarization transfer [165,166,167]. As the spin diffusion period tSD increases, polarization is then transferred to adjacent, less dynamic domains, enabling the identification of molecular domains based on their spatial proximity and mobility. This is demonstrated in Fig. 13d for water-hydrated switchgrass at natural isotopic abundance: as the spin diffusion period is increased, magnetization that originated on from water penetrates first into water-hydrated regions and then into water-deficient areas, like cellulose microfibrils, and spin-diffusion buildup curves are obtained. This method provides insights into the microstructural organization and relative sizes of different regions within biomaterials [168].
Variable contact time CP-MAS
The variable contact time (VCT) experiment represents a very basic pseudo-2D approach in ssNMR, in which the time-dependent dynamics of the polarization transfer from abundant nuclei like 1H to less sensitive ones such as 13C is monitored. Variation of the contact time in a CP pulse sequence results in a buildup then decay of 13C signals corresponding, respectively, to the TCH constant and T1ρ(H) rotating frame relaxation time for each resolved carbon. Interpretation of both TCH and T1ρ can yield important insights into the efficiency of polarization transfer, the strength of dipolar couplings, inter-nuclear distances, and local molecular dynamics. The shorter the TCH, the stronger the dipolar interactions and the more effective the transfer [169]. Kobayashi et al. used TCH measurements to show that ether linkages between syringyl and guaiacyl units are broken and new linkages formed, reflecting the recondensation of lignin, especially for dilute H2SO4 treatment at higher temperatures for powdered poplar wood [161]. Their data revealed a small rise in TCH values with time at the 120 and 130 °C treatments, indicating a slight increase in atomic-level mobility. In contrast, the treatments at 140 and 150 °C revealed a slight decrease in TCH values with time that reflected reduced atomic-level mobility during these conditions. In a study by Santoni et al. [151] TCH measurements were applied to untreated wood of Picea abies from Finland, Poland, and Italy. The study found that the chemical and physical properties of the wood varied by provenance, allowing classification based on growth conditions. Although no consistent trend in TCH values was identified across the wood samples, origin-dependent differences in T1ρ(H) relaxation times were observed. Specifically, the Finnish samples exhibited shorter T1ρ(H) times, possibly indicating more dynamic molecular structures, which may reflect the local climatic conditions and environmental influences on wood properties.
Detect strength of heteronuclear dipolar couplings
Another valuable pseudo-2D technique for elucidating the motional amplitudes of lignocellulose, measured using motionally averaged 13C–1H dipolar couplings, is the dipolar DIPSHIFT experiment (Fig. 13e) [170]. The 1H–13C DIPSHIFT is a type of separated local field (SLF) ssNMR experiment aimed at correlating chemical shifts with the strength of the heteronuclear dipolar coupling interactions [171, 172]. In this experiment, the initial 13C magnetization is generated via cross-polarization, followed by a variable evolution time t1 during which the 13C magnetization evolves under the 1H–13C dipolar interaction. Throughout this period, 1H–1H homonuclear interactions are eliminated using homonuclear decoupling sequences such as frequency-switched Lee–Goldburg (FSLG) and MREV-8 [173, 174]. The chemical shift is then refocused using a π pulse placed after one rotor period; as a result, the 13C spins evolve exclusively under the 1H–13C dipolar coupling. The resulting spectrum displays 13C isotropic chemical shifts in the direct dimension and 1H–13C dipolar dephasing curves in the indirect dimension, which can be fitted to extract the apparent 1H–13C dipolar couplings. Dipolar curves are generated for one rotor period, and the strength of the 1H–13C coupling is reflected in the modulation depth of the curve. In fully rigid systems, the strength of the 1H–13C coupling remains unaffected by molecular averaging, leading to the maximum modulation depth of the DIPSHIFT curve at half the rotor period. However, when dynamic processes emerge in the intermediate regime, the effective 1H–13C coupling interaction diminishes, resulting in reduced depth of the DIPSHIFT curve. By comparing the observed dipolar coupling strengths to those from a fully rigid system (23 kHz for a rigid 1.1 Å aliphatic C–H bond), an order parameter S can be derived, representing the degree of reduction in dipolar coupling due to intermediate molecular motion. The apparent coupling is then divided by the scaling factors associated with the homonuclear decoupling sequences—0.577 for FSLG and 0.47 for MREV-8—to obtain the true coupling.
Zhao et al. demonstrated the utility of dynamic NMR studies on unlabeled rice (Oryza sativa) materials, rich in cellulose, hemicelluloses (primarily xylan), and lignin, to assess their native states [47]. Their work effectively combined room-temperature measurements of polymer dynamics with DNP-enhanced natural-abundance 2D 13C-13C correlation experiments, facilitating efficient screening of a diverse array of lignocellulosic materials found in nature or engineered in vitro. Employing the DIPSHIFT experiment, they investigated the CH dipolar coupling strengths between wild-type rice and double mutants of the homolog CTL2, which is implicated in cellulose biosynthesis regulation. Notably, their data revealed that all polysaccharides exhibited high rigidity, characterized by large order parameters exceeding 0.90 (Fig. 14a).
2D ssNMR unveils the complex structure of natural abundance biomass. a Simulated dipolar curves for CH and CH2 indicate increased dipolar coupling at greater depths. The initial slice of the pseudo-2D DIPSHIFT experiment on rice cell walls revealed distinct cellulose and xylan signals, with elevated order parameters in dipolar curves for both the wild type and double mutant of interior cellulose carbon 6 (i6). Panel adapted with permission of ref [47]. b Narrow line shaped in 1H shows cellulose increased mobility in acid-treated poplar. Panel adapted with permission of ref [175]. c Utilize 2D WISE to elucidate the relationship between crystalline and amorphous cellulose domains and the associated water pool. Panel adapted with permission of ref [176]. d Identification of functional groups in enzymatic hydrolyzed corn stover using 1H–13C FSLG HETCOR without spin diffusion. Panel adapted with permission of ref [144]. e Interactions between archeological wood and consolidant (Kauramin) [177]. f At 300 °C, the acetate groups in lignin and hemicelluloses exhibit incomplete conversion, attributed to the constraints imposed by the macromolecular network present in native Miscanthus. Panel adapted with permission of ref [178]
2D (13C–1H) wideline separation (WISE)
Another form of separated local field ssNMR technique is 2D wide-line separation (WISE) (Fig. 13e) [145, 179], which is designed to correlate biopolymer structure/conformation in the direct dimension (13C chemical shift) with mobility and dynamics via the 1H linewidth in the indirect dimension. The 1H wideline spectrum is a reflection of 1H–1H dipolar couplings of the protons in the proximity of the observed 13C site, and thus is a proxy for molecular dynamics. A broad (~ 50–70 kHz FWHM) 1H profile signifies a rigid proton system with strong homonuclear dipolar couplings, while narrowing of the 1H wideline spectrum reveals averaging of the dipolar coupling network due to segmental motion. Assuming site-specific information is requested, short cross-polarization contact times are usually employed to select 1H–13C spin pairs and limit the effects of spin-diffusion and averaging of dynamics environments [123]. 2D 13C–1H WISE spectra (without spin-diffusion) on native poplar wood was compared to poplar wood/silica composites prepared under basic and acidic conditions [175]. 1H wideline traces associated with bulk structural polysaccharides (73 ppm) revealed broad ~ 62 kHz profiles for native and wood–silica composites prepared under basic conditions, but under acidic conditions a narrow ~ 10 kHz FWHM component is superimposed with the broad profile (Fig. 14b). This observation of a subfraction of dynamic cellulose in the acid-prepared wood/silica composite material was correlated with a removal of lignin as observed by CP-MAS, and was attributed to a partial disruption of cellulose and better penetration of silica precursors into the cell wall interior. 2D WISE was also used to investigate 1H lineshape changes to Whatman filter paper cellulose under differing moisture states. Results showed narrowing of the 1H profiles for amorphous/surface cellulose signals (84 ppm) with increased humidity, but no narrowing is observed for crystalline/interior signals (89 ppm) [149].
A proton spin-diffusion element added to the 2D WISE experiment allows one to monitor the time-dependent equilibration of phase-separated 1H spin baths of differing dynamics, which can yield morphology or domain-size information [145]. For example, Capitani et al. [176] used 2D WISE with spin-diffusion to reveal the spatial distribution of nanoporous water in paper cellulose. The authors leverage the dynamics difference between water and cellulose; protons from water have a narrow 1H spectrum of ~ 1.5 kHz while protons associated with cellulose are ~ 62 kHz FWHM (Fig. 14c). At short spin-diffusion mixing times the narrow 1H profile of water is not visible in the WISE spectrum at 13C sites associated with cellulose, which instead show characteristically broad 1H wideline profiles. But as the mixing time increases, 1H magnetization from the mobile water regions transfers to the cellulose and is readable as a narrow 1H component in the 2D WISE spectrum. Tracking the time-dependent equilibration of mobile and rigid proton spin baths ultimately led to an estimate of ~ 3 nm separation between crystallite cellulose cores within a water-rich amorphous matrix. 2D WISE has also used to investigate spatial proximities and local dynamics of suberin in potato skins [180, 181], cutin in tomato skins [182], and to probe hydration effects of polysaccharides within onion primary cell walls [150].
2D (13C–1H) heteronuclear correlation (HETCOR)
Arguably the typical NMR method for whole biomass analysis is whole-cell gel-state HSQC NMR pioneered by the John Ralph lab, in which biomass is pulverized (ball-milled) and swollen in a plasticizing NMR solvent like d6-DMSO and analyzed by HSQC NMR to reveal directly bonded 1H–13C pairs [65]. Although the resolution-enhancement is impressive, limitations include only protonated carbons detected, physical modifications to the cell wall caused by mechanical milling, sample heating, as well as issues like heterogeneous solvent penetration, leaving room for 1H–13C correlation experiments of unaltered materials in the solid state. In the native solid-state however, observing correlation of 13C with associated 1H nuclei is severely hindered by strong homonuclear dipolar couplings in a rigid proton network. This is exemplified in the above-discussed 2D WISE experiment, which correlates 13C with associated protons but chemical shift information is impossible to interpret due to the breadth of the 1H profile. One can partially overcome these issues by implementing homonuclear decoupling methods like FLSG decoupling where during 1H chemical shift evolution 1H–1H homonuclear dipole–dipole interactions are thereby suppressed (Fig. 13e) [173, 183]. This is the basis of the 2D 1H–13C FSLG-HETCOR, which correlates the chemical shifts of 13C with proximal 1H in the solid state [184]. Like the 2D WISE experiment, 2D 1H–13C HETCOR can be collected without spin-diffusion for site-specific information or with an added 1H spin-diffusion block to probe the spatial proximities of differing environments.
In the same study highlighted previously for spectral editing ssNMR tools, Mao and colleagues applied 1H–13C FSLG-HETCOR on corn stover residuals after DAP and enzymatic hydrolysis [144]. These data revealed that cellulose and/or hemicellulose within the corn stover residue after conversion is tightly associated with lignin, as evident by the observed correlation between 13C of 89 and 105 ppm with aromatic protons in the 7 ppm range [144]. Adding a dipolar dephasing element to the 2D HETCOR also helped correlate non-trivial unprotonated aromatic signals with nearby non-bonded protons (Fig. 14d).
In a related study, 2D 1H–13C HETCOR was used to compare residual enzymatic lignin (REL) and traditional milled wood lignin (MWL) from Loblolly pine [185]. Here, a 1H spin-diffusion element was included in the 2D HETCOR, which revealed REL is more heterogeneous compared to MWL. In a unique example, 2D FSLG-HETCOR ssNMR was used to show resin/lignin interactions within a roman warship wooden archeological artifact after consolidation (Fig. 14e) [177]. In particular, 1H–13C correlations were observed between Kauramin carbons and lignin aromatic protons, providing direct evidence of pi–pi stacking between lignin and the triazine ring of the Kauramin binder. 2D 1H–13C FSLG HETCOR ssNMR is also highlighted as a key tool for characterizing charred organic materials and revealing the mechanisms at play during biochar production, in particular because solid materials are analyzed in their unaltered state without any sample preparation needs (Fig. 14f) [178, 186, 187]. Brech et al. demonstrated that below 300 °C, xylan is the primary source of aromatic compounds, whereas above this temperature, cellulose contributes to aromatization in silvergrass (Miscanthus). This transition is indicated by the cross peaks observed at 173–172/3.7–3.1 ppm in 1H–13C HETCOR, suggesting interactions between lignin and carbohydrates within the biomass network. For an expanded reading on utilizing 2D HETCOR for understanding cellulose-based pyrolysis residues, please see the tutorial article by Knicker.
Future directions: fast MAS, 1H detection, dynamic nuclear polarization (DNP)
On the bleeding edge of natural abundant ssNMR capabilities include ultrafast MAS and DNP, both requiring specialized hardware. Ultrafast MAS ssNMR improves the spectral resolution by spinning samples with high frequencies (40–150 kHz), which effectively averages out anisotropic interactions such as dipolar couplings and chemical shift anisotropy [188]. In that way, proton–proton dipolar couplings—the major source of broadening—are eliminated (Fig. 15a), improving sensitivity and clarity, especially for the low-abundance nuclei like 13C and 15N. Fast MAS also suppresses the broadening due to quadrupolar interactions in nuclei such as 14N and 17O, making this technique particularly suitable for high-resolution work on biomass, proteins, nucleic acids, and materials. This allows much better quantification and structural work in natural abundance samples, although this comes with serious drawbacks of reduced sample volume from smaller rotors. Direct measurement of 1H spectra from rigid organic solids also enables 1H-detected 2D correlation experiments which bring substantially improved sensitivity. In a recent first, Yuan et al. utilized ultrafast MAS at 150 kHz combined with high magnetic field (800 MHz) NMR to investigate spruce and maple wood as well as cotton cellulose [105, 187]. This approach enabled 1H-detected 2D 1H–13C HETCOR experiments at natural abundance for a detailed analysis of cellulose conformations in native and heat-treated wood samples (Fig. 15b). The improved resolution, particularly in the 1H dimension, distinctly revealed three cellulose conformations assigned to tg, gt, and gg. Results demonstrated that softwoods like spruce exhibit higher packing order and crystallinity compared to hardwoods like maple. Furthermore, thermal treatments were shown to result in hemicellulose loss, increased cellulose disorder, and greater packing heterogeneity. As ultrafast MAS hardware becomes more prevalent, similar applications to understanding lignocellulose and its deconstruction can be expected.
Future directions in advanced natural abundance biomass analysis. a 1H detection in solids achieved through ultrafast MAS, with resolution improving at higher spinning speeds. Panel adapted with permission of ref [188]. b 1H{13C} CP-HETCOR spectra of spruce at varying spinning frequencies and fields show improved resolution with higher spinning rates and field strengths. Panel adapted with permission of ref [189]. c Chemical structures of symmetric and asymmetric biradicals employed in DNP polarizing agents. d DNP sensitivity enhancement through polarization transfer via biradicals. Panel adapted with permission of ref [47]. e DNP-enhanced 2D 13C–13C refocused-INADEQUATE of biomass and resistant starch at natural abundance. Panel adapted with permission of ref [116]
DNP has revolutionized NMR spectroscopy by significantly enhancing the sensitivity of measurements, especially for natural abundance biomass materials with inherently low signal intensities. By leveraging biradicals such as AMUPol and AsymPolPOk (Fig. 15c), DNP dramatically amplifies signal intensity, reducing acquisition times by orders of magnitude and substantially improving signal-to-noise ratios [190]. This enhancement makes DNP an indispensable tool for studying complex systems and heterogeneous materials. DNP is particularly well-suited for acquiring high-resolution 13C–13C correlation spectra, which are essential for unraveling structural details in challenging systems like lignocellulosic biomass. It enables a broad range of advanced NMR experiments, including single quantum coherence transfer methods such as PDSD, CORD, and DARR, as well as double quantum coherence transfer techniques like INADEQUATE, facilitating in-depth structural characterization with unparalleled efficiency. Innovative developments have further extended the utility of DNP beyond its conventional framework. A major breakthrough came from the use of a sorbitol-based glass matrix, which enables DNP experiments to be conducted at elevated temperatures above 200 K [191]. This advancement broadens the range of experimental conditions, providing enhanced flexibility and enabling studies under more biologically or industrially relevant conditions. Another milestone in DNP is the introduction of matrix-free protocols, which address longstanding challenges related to the homogeneity of traditional DNP matrices, typically mixtures of d₈-glycerol, D2O, and H2O. Zhao et al. demonstrated that a matrix-free approach using minimal D2O achieved an unprecedented enhancement factor (ε) of 57 (Fig. 15d), corresponding to a timesaving of 3249-fold. This approach mitigates issues such as aggregation and phase separation of glass-forming solvents at cryogenic temperatures, ensuring superior sensitivity, resolution, and reproducibility. These advancements simplify the experimental setup while significantly lowering operational complexity and cost, making DNP a more accessible and versatile technique for studying lignocellulosic components. In another example, Sluiter et al. confirm selective starch removal in biomass, with DNP-enhanced 13C INADEQUATE showing the disappearance of starch signals (~ 100–102 ppm) while retaining characteristic cellulose peaks (89 and 105 ppm) (Fig. 15e) [116]. The observed 101–102 ppm single quantum and 176 ppm double quantum correlation matches threefold helical screw xylan, further validating the method’s accuracy in distinguishing cellulosic glucan from starch.
Conclusions and perspectives
Solid-state NMR has emerged as an indispensable tool for understanding the intricate molecular structure and dynamics of lignocellulosic biomass and related bioenergy materials. Its unique capability to provide non-destructive, high-resolution insights into the composition, morphology, and transformations of rigid solids at natural abundance sets it apart from conventional analytical techniques. Throughout this review we cover common and routine ssNMR methods in addition to more advanced techniques including spectral editing, accessing dynamics-based information, and select 2D ssNMR methods practical at natural abundance. Our goal is to provide fundamental information and highlight the application of ssNMR methods at natural abundance to understand lignocellulosic biomass composition, structure and conversion and for bioenergy and carbon management applications. Solid-state NMR data provide essential insights of lignocellulosic biomass that unite our fundamental understanding of plant cell wall synthesis and architecture, as well as its deconstruction, decay and conversion to solids and liquids relevant in the bioeconomy. Further, the frontier of using natural abundance ssNMR for applications in carbon management and a sustainable economy is immense. Conversion and deconstruction kinetics, biosynthesis processes, system-scale ecosystem dynamics among lignocellulose and microbes, as well as biomass and biochar durability related to long-term carbon sequestration and storage could benefit from the information obtainable by ssNMR.
A central theme of this review is the unparalleled advantage of ssNMR in maintaining the integrity of the sample during analysis, preserving its natural state and complex structure. This advantage is particularly crucial in bioenergy research, where understanding the native organization of plant cell walls, the interactions between cellulose, hemicellulose, and lignin, and the changes induced during biotic and abiotic conversion processes are key to improving efficiency. The ability to observe these materials in their unaltered state has revealed fundamental insights into the mechanisms of biomass deconstruction, guiding the development of more effective pretreatment and conversion strategies.
Despite its numerous advantages, solid-state NMR also presents notable challenges and limitations. One primary drawback is its inherently low sensitivity, particularly for nuclei like 13C at natural abundance, which necessitates long acquisition times and can limit throughput. Additionally, the cost and complexity of high-field ssNMR instrumentation, coupled with the need for specialized expertise in experimental setup and data interpretation, can be prohibitive for some research groups. Another limitation is the difficulty in resolving overlapping signals in complex mixtures, which can obscure detailed compositional or structural insights without the use of advanced methods such as multi-dimensional NMR or isotopic labeling. Finally, ssNMR data acquisition is resource-intensive, requiring significant time and computational efforts, which may constrain its routine application in large-scale or high-throughput studies. Addressing these challenges through technological advancements, such as faster spinning rates, DNP, and artificial intelligence-driven data analysis, could expand the accessibility and applicability of ssNMR.
In conclusion, ssNMR not only enhances our fundamental understanding of biomass, but also provides critical data to inform bioenergy solutions and carbon-efficient practices. As the global community intensifies efforts to achieve sustainable energy solutions, ssNMR stands poised to play a pivotal role. Its application extends beyond lignocellulosic biomass to include biochar, carbon sequestration materials, and other bio-derived polymers, underscoring its versatility in addressing carbon management challenges.
Availability of data and materials
The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan). Data is available in the manuscript and any data not provided is available upon request from the coauthors. No datasets were generated or analysed during the current study.
Abbreviations
- ssNMR:
-
Solid-state nuclear magnetic resonance
- 1D:
-
One dimensional
- 2D:
-
Two dimensional
- 3D:
-
Three dimensional
- CBP:
-
Consolidated bioprocessing
- NREL:
-
National Renewable Energy Laboratory
- DMR:
-
Deacetylation and mechanical refining
- RCF:
-
Reductive catalytic fractionation
- HPLC:
-
High-pressure liquid chromatography
- GC:
-
Gas chromatography
- FTIR:
-
Fourier transform infrared
- UV–Vis:
-
Ultraviolet–visible
- py-MBMS:
-
Pyrolysis-molecular beam mass spectroscopy
- NMR:
-
Nuclear magnetic resonance
- SEM:
-
Scanning electron microscopy
- TEM:
-
Transmission electron microscopy
- TGA:
-
Thermogravimetric analysis
- DSC:
-
Differential scanning calorimetry
- XRD:
-
X-ray diffraction
- LAP:
-
Laboratory analytical procedures
- µ :
-
Magnetic moment
- µ 0 :
-
Net magnetic moment
- M 0 :
-
Net magnetization
- N :
-
Number of detectable spins
- B 0 :
-
External magnetic field
- γ :
-
Gyromagnetic ratio
- I :
-
Spin quantum number
- DP:
-
Direct polarization
- CP:
-
Cross-polarization
- HH:
-
Hartmann–Hahn
- VACP:
-
Variable-amplitude cross-polarization
- RAMP-CP:
-
Ramped-amplitude cross-polarization sequence
- DNP:
-
Dynamic nuclear polarization
- MAS:
-
Magic angle spinning
- rf:
-
Radiofrequency
- B 1 :
-
Field strength
- J :
-
Scalar coupling
- CS:
-
Chemical shielding
- Q :
-
Quadrupolar interactions
- σ :
-
Shielding constant
- δ :
-
Chemical shifts
- ppm:
-
Parts per million
- TMS:
-
Tetramethylsilane
- DSS:
-
Sodium trimethylsilyl propane sulfonate
- TOSS:
-
TOtal Suppression of Spinning Sidebands
- d1:
-
Recycle delay
- FID:
-
Free induction decay
- GAX:
-
Glucuronoarabinoxylan
- GlcA:
-
Glucuronic acid
- GX:
-
Glucuronoxylan
- HG:
-
Homogalacturonan
- RG:
-
Rhamnogalacturonan
- G:
-
Guaiacyl
- S:
-
Syringyl
- H:
-
p-Hydroxyphenyl
- TMSP:
-
Sodium-3-trimethylsilylpropionate
- SB:
-
Sugarcane bagasse
- AH:
-
Acid hydrolysis
- HSQC:
-
Heteronuclear single quantum coherence
- CrI:
-
Cellulose crystallinity index
- CI:
-
Crystallinity index
- ORNL:
-
Oak Ridge National Laboratory
- ILs:
-
Ionic liquids
- EimCl:
-
Ethylimidazolium chloride
- HPAC:
-
Hydrogen peroxide-acetic acid
- COSLIF:
-
Cellulose solvent- and organic solvent-based lignocellulose fractionation
- TCS:
-
Total carbohydrate solubilization
- HTT:
-
Heat treatment temperatures
- CSA:
-
Chemical shift anisotropy
- PSRE:
-
Proton Spin Relaxation Editing
- LG:
-
Lee–Goldburg
- kHz:
-
Kilo hertz
- MHz:
-
Mega hertz
- DD:
-
Dipolar dephasing
- DAP:
-
Dilute acid pretreatment
- PDSD:
-
Proton driven spin diffusion
- PCA:
-
Principal component analysis
- HWP:
-
Hot water pretreatment
- CAD:
-
Cinnamyl alcohol dehydrogenase
- DIPSHIFT:
-
Dipolar chemical shift correlation
- WISE:
-
Wideline separation
- HETCOR:
-
Heteronuclear correlation
- SWG:
-
Switchgrass
- VCT:
-
Variable contact time
- SLF:
-
Separated local field
- FSLG:
-
Frequency switched Lee–Goldburg
- S :
-
Order parameter
- FWHM:
-
Half width half maximum
- REL:
-
Residual enzymatic lignin
- MWL:
-
Milled wood lignin
References
Guragain YN, Vadlani PV. Renewable biomass utilization: a way forward to establish sustainable chemical and processing industries. Clean Technol. 2021;3:243–59.
Singh N, Singhania RR, Nigam PS, Dong C-D, Patel AK, Puri M. Global status of lignocellulosic biorefinery: challenges and perspectives. Biores Technol. 2022;344: 126415.
Mujtaba M, Fernandes Fraceto L, Fazeli M, Mukherjee S, Savassa SM, Araujo De Medeiros G, et al. Lignocellulosic biomass from agricultural waste to the circular economy: a review with focus on biofuels, biocomposites and bioplastics. J Clean Prod. 2023;402: 136815.
Alzagameem A, El Khaldi-Hansen B, Kamm B, Schulze M. Lignocellulosic biomass for energy, biofuels, biomaterials, and chemicals. In: Vaz S Jr, editor. Biomass and green chemistry: building a renewable pathway. Cham: Springer International Publishing; 2018. p. 95–132. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/978-3-319-66736-2_5.
Naik SN, Goud VV, Rout PK, Dalai AK. Production of first and second generation biofuels: a comprehensive review. Renew Sustain Energy Rev. 2010;14:578–97.
Sims REH, Mabee W, Saddler JN, Taylor M. An overview of second generation biofuel technologies. Biores Technol. 2010;101:1570–80.
Menon V, Rao M. Trends in bioconversion of lignocellulose: Biofuels, platform chemicals & biorefinery concept. Prog Energy Combust Sci. 2012;38:522–50.
Kumar AK, Sharma S. Recent updates on different methods of pretreatment of lignocellulosic feedstocks: a review. Bioresour Bioprocess. 2017;4:7.
Shukla A, Kumar D, Girdhar M, Kumar A, Goyal A, Malik T, et al. Strategies of pretreatment of feedstocks for optimized bioethanol production: distinct and integrated approaches. Biotechnol Biofuels. 2023;16:44.
Lynd LR, van Zyl WH, McBride JE, Laser M. Consolidated bioprocessing of cellulosic biomass: an update. Curr Opin Biotechnol. 2005;16:577–83.
Li Z, Waghmare PR, Dijkhuizen L, Meng X, Liu W. Research advances on the consolidated bioprocessing of lignocellulosic biomass. Eng Microbiol. 2024;4: 100139.
Balch ML, Holwerda EK, Davis MF, Sykes RW, Happs RM, Kumar R, et al. Lignocellulose fermentation and residual solids characterization for senescent switchgrass fermentation by Clostridium thermocellum in the presence and absence of continuous in situ ball-milling. Energy Environ Sci. 2017;10:1252–61.
Holwerda EK, Worthen RS, Kothari N, Lasky RC, Davison BH, Fu C, et al. Multiple levers for overcoming the recalcitrance of lignocellulosic biomass. Biotechnol Biofuels. 2019;12:15.
Paye JMD, Guseva A, Hammer SK, Gjersing E, Davis MF, Davison BH, et al. Biological lignocellulose solubilization: comparative evaluation of biocatalysts and enhancement via cotreatment. Biotechnol Biofuels. 2016;9:8.
Chen X, Shekiro J, Pschorn T, Sabourin M, Tao L, Elander R, et al. A highly efficient dilute alkali deacetylation and mechanical (disc) refining process for the conversion of renewable biomass to lower cost sugars. Biotechnol Biofuels. 2014;7:98.
Chen X, Kuhn E, Jennings EW, Nelson R, Tao L, Zhang M, et al. DMR (deacetylation and mechanical refining) processing of corn stover achieves high monomeric sugar concentrations (230 g L−1) during enzymatic hydrolysis and high ethanol concentrations (>10% v/v) during fermentation without hydrolysate purification or concentration. Energy Environ Sci. 2016;9:1237–45.
Zhang L, Xu C, Champagne P. Overview of recent advances in thermo-chemical conversion of biomass. Energy Convers Manag. 2010;51:969–82.
Chiaramonti D, Lehmann J, Berruti F, Giudicianni P, Sanei H, Masek O. Biochar is a long-lived form of carbon removal, making evidence-based CDR projects possible. Biochar. 2024;6:81.
Ragauskas AJ, Beckham GT, Biddy MJ, Chandra R, Chen F, Davis MF, et al. Lignin valorization: improving lignin processing in the biorefinery. Science. 2014;344:1246843.
Beckham GT, Johnson CW, Karp EM, Salvachúa D, Vardon DR. Opportunities and challenges in biological lignin valorization. Curr Opin Biotechnol. 2016;42:40–53.
Sethupathy S, Murillo Morales G, Gao L, Wang H, Yang B, Jiang J, et al. Lignin valorization: status, challenges and opportunities. Biores Technol. 2022;347: 126696.
Hoque M, Kamal S, Raghunath S, Foster EJ. Unraveling lignin degradation in fibre cement via multidimensional fluorometry. Sci Rep. 2023;13:8385.
Renders T, Van Den Bosch S, Koelewijn S-F, Schutyser W, Sels BF. Lignin-first biomass fractionation: the advent of active stabilisation strategies. Energy Environ Sci. 2017;10:1551–7.
Abu-Omar MM, Barta K, Beckham GT, Luterbacher JS, Ralph J, Rinaldi R, et al. Guidelines for performing lignin-first biorefining. Energy Environ Sci. 2021;14:262–92.
Anderson EM, Katahira R, Reed M, Resch MG, Karp EM, Beckham GT, et al. Reductive catalytic fractionation of corn stover lignin. ACS Sustainable Chem Eng. 2016;4:6940–50.
Renders T, Van Den Bossche G, Vangeel T, Van Aelst K, Sels B. Reductive catalytic fractionation: state of the art of the lignin-first biorefinery. Curr Opin Biotechnol. 2019;56:193–201.
Sluiter JB, Ruiz RO, Scarlata CJ, Sluiter AD, Templeton DW. Compositional analysis of lignocellulosic feedstocks. 1. Review and description of methods. J Agric Food Chem. 2010;58:9043–53.
Cai J, He Y, Yu X, Banks SW, Yang Y, Zhang X, et al. Review of physicochemical properties and analytical characterization of lignocellulosic biomass. Renew Sustain Energy Rev. 2017;76:309–22.
Singh SS, Lim L-T, Manickavasagan A. Imaging and spectroscopic techniques for microstructural and compositional analysis of lignocellulosic materials: a review. Biomass Conv Bioref. 2023;13:499–517.
Dupree R, Simmons TJ, Mortimer JC, Patel D, Iuga D, Brown SP, et al. Probing the molecular architecture of arabidopsis thaliana secondary cell walls using two- and three-dimensional 13C solid state nuclear magnetic resonance spectroscopy. Biochemistry. 2015;54:2335–45.
Simmons TJ, Mortimer JC, Bernardinelli OD, Pöppler A-C, Brown SP, deAzevedo ER, et al. Folding of xylan onto cellulose fibrils in plant cell walls revealed by solid-state NMR. Nat Commun. 2016;7:13902.
Grantham NJ, Wurman-Rodrich J, Terrett OM, Lyczakowski JJ, Stott K, Iuga D, et al. An even pattern of xylan substitution is critical for interaction with cellulose in plant cell walls. Nat Plants. 2017;3:859–65.
Wierzbicki MP, Maloney V, Mizrachi E, Myburg AA. Xylan in the middle: understanding xylan biosynthesis and its metabolic dependencies toward improving wood fiber for industrial processing. Front Plant Sci. 2019;10:1–29.
Kang X, Kirui A, Dickwella Widanage MC, Mentink-Vigier F, Cosgrove DJ, Wang T. Lignin-polysaccharide interactions in plant secondary cell walls revealed by solid-state NMR. Nat Commun. 2019;10:347.
Terrett OM, Lyczakowski JJ, Yu L, Iuga D, Franks WT, Brown SP, et al. Molecular architecture of softwood revealed by solid-state NMR. Nat Commun. 2019;10:4978.
Addison B, Stengel D, Bharadwaj VS, Happs RM, Doeppke C, Wang T, et al. Selective one-dimensional 13 C– 13 C spin-diffusion solid-state nuclear magnetic resonance methods to probe spatial arrangements in biopolymers including plant cell walls, peptides, and spider silk. J Phys Chem B. 2020;124:9870–83.
Gao Y, Lipton AS, Wittmer Y, Murray DT, Mortimer JC. A grass-specific cellulose–xylan interaction dominates in sorghum secondary cell walls. Nat Commun. 2020;11:1–10.
Duan P, Kaser SJ, Lyczakowski JJ, Phyo P, Tryfona T, Dupree P, et al. Xylan structure and dynamics in native brachypodium grass cell walls investigated by solid-state NMR spectroscopy. ACS Omega. 2021;6:15460–71.
Munson CR, Gao Y, Mortimer JC, Murray DT. Solid-state nuclear magnetic resonance as a tool to probe the impact of mechanical preprocessing on the structure and arrangement of plant cell wall polymers. Front Plant Sci. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fpls.2021.766506.
Kirui A, Zhao W, Deligey F, Yang H, Kang X, Mentink-Vigier F, et al. Carbohydrate-aromatic interface and molecular architecture of lignocellulose. Nat Commun. 2022;13:538.
Addison B, Bu L, Bharadwaj V, Crowley MF, Harman-Ware AE, Crowley MF, et al. Atomistic, macromolecular model of the Populus secondary cell wall informed by solid-state NMR. Sci Adv. 2024;10:eadi7965.
Foston M. Advances in solid-state NMR of cellulose. Curr Opin Biotechnol. 2014;27:176–84.
Baccile N, Falco C, Titirici M-M. Characterization of biomass and its derived char using 13C-solid state nuclear magnetic resonance. Green Chem. 2014;16:4839–69.
Mao J, Cao X, Chen N. Characterization of biochars using advanced solid-state 13C nuclear magnetic resonance spectroscopy. In: Lee JW, editor. Advanced biofuels and bioproducts. New York: Springer; 2013. p. 47–55. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/978-1-4614-3348-4_5.
Mao J, Cao X, Olk DC, Chu W, Schmidt-Rohr K. Advanced solid-state NMR spectroscopy of natural organic matter. Prog Nucl Magn Reson Spectrosc. 2017;100:17–51.
El Hariri El Nokab M, van der Wel PCA. Use of solid-state NMR spectroscopy for investigating polysaccharide-based hydrogels: a review. Carbohydr Polym. 2020;240:1–9.
Zhao W, Kirui A, Deligey F, Mentink-Vigier F, Zhou Y, Zhang B, et al. Solid-state NMR of unlabeled plant cell walls: high-resolution structural analysis without isotopic enrichment. Biotechnol Biofuels. 2021;14:1–14.
Schaefer J, Stejskal EO. Carbon-13 nuclear magnetic resonance of polymers spinning at the magic angle. J Am Chem Soc. 1976;98:1031–2.
Stejskal EO, Schaefer J, Waugh JS. Magic-angle spinning and polarization transfer in proton-enhanced NMR. J Magn Reson. 1969;1977(28):105–12.
Hartmann SR, Hahn EL. Nuclear double resonance in the rotating frame. Phys Rev. 1962;128:2042–53.
Barbara TM, Williams EH. Modulated sequences for cross polarization during high-speed MAS. J Magn Reson. 1969;1992(99):439–42.
Metz G, Wu XL, Smith SO. Ramped-amplitude cross polarization in magic-angle-spinning NMR. J Magn Reson, Ser A. 1994;110:219–27.
Bloch F. Theory of line narrowing by double-frequency irradiation. Phys Rev. 1958;111:841–53.
Vanderhart DL, Earl WL, Garroway AN. Resolution in 13C NMR of organic solids using high-power proton decoupling and magic-angle sample spinning. J Magn Reson. 1969;1981(44):361–401.
Johnson RL, Schmidt-Rohr K. Quantitative solid-state 13C NMR with signal enhancement by multiple cross polarization. J Magn Reson. 2014;239:44–9.
Dixon WT, Schaefer J, Sefcik MD, Stejskal EO, McKay RA. Total suppression of sidebands in CPMAS C-13 NMR. J Magn Reson. 1969;1982(49):341–5.
Caffall KH, Mohnen D. The structure, function, and biosynthesis of plant cell wall pectic polysaccharides. Carbohyd Res. 2009;344:1879–900.
Yang H, Benatti MR, Karve RA, Fox A, Meilan R, Carpita NC, et al. Rhamnogalacturonan-I is a determinant of cell–cell adhesion in poplar wood. Plant Biotechnol J. 2020;18:1027–40.
Updegraff DM. Semimicro determination of cellulose in biological materials. Anal Biochem. 1969;32:420–4.
Klemm D, Heublein B, Fink H, Bohn A. Cellulose: fascinating biopolymer and sustainable raw material. Angew Chem Int Ed. 2005;44:3358–93.
Zhang B, Gao Y, Zhang L, Zhou Y. The plant cell wall: biosynthesis, construction, and functions. JIPB. 2021;63:251–72.
Rosini E, Molinari F, Miani D, Pollegioni L. Lignin valorization: production of high value-added compounds by engineered microorganisms. Catalysts. 2023;13:555.
Gil AM, Lopes M, Rocha J, Pascoal NC. A 13C solid state nuclear magnetic resonance spectroscopic study of cork cell wall structure: the effect of suberin removal. Int J Biol Macromol. 1997;20:293–305.
Kang X, Zhao W, Dickwella Widanage MC, Kirui A, Ozdenvar U, Wang T. CCMRD: a solid-state NMR database for complex carbohydrates. J Biomol NMR. 2020;74:239–45.
Mansfield SD, Kim H, Lu F, Ralph J. Whole plant cell wall characterization using solution-state 2D NMR. Nat Protoc. 2012;7:1579–89.
Ralph SA, Ralph J, Lu F. NMR database of lignin and cell wall model compounds. United States; 1AD. https://www.osti.gov/biblio/2409191
Scheller HV, Ulvskov P. Hemicelluloses. Annu Rev Plant Biol. 2010;61:263–89.
Curry TM, Peña MJ, Urbanowicz BR. An update on xylan structure, biosynthesis, and potential commercial applications. The Cell Surface. 2023;9: 100101.
Mohnen D. Pectin structure and biosynthesis. Curr Opin Plant Biol. 2008;11:266–77.
Boerjan W, Ralph J, Baucher M. Lignin biosynthesis. Annu Rev Plant Biol. 2003;54:519–46.
Ralph J, Lapierre C, Boerjan W. Lignin structure and its engineering. Curr Opin Biotechnol. 2019;56:240–9.
Ralph J. Hydroxycinnamates in lignification. Phytochem Rev. 2010;9:65–83.
Boeriu CG, Bravo D, Gosselink RJA, Van Dam JEG. Characterisation of structure-dependent functional properties of lignin with infrared spectroscopy. Ind Crops Prod. 2004;20:205–18.
Pereira H. Chemical composition and variability of cork from Quercus suber L. Wood SciTechnol. 1988;22:211–8.
Graça J. Suberin: the biopolyester at the frontier of plants. Front Chem. 2015;3:11.
Harman-Ware AE, Sparks S, Addison B, Kalluri UC. Importance of suberin biopolymer in plant function, contributions to soil organic carbon and in the production of bio-derived energy and materials. Biotechnol Biofuels. 2021;14:1–21.
Woolfson KN, Esfandiari M, Bernards MA. Suberin biosynthesis, assembly, and regulation. Plants. 2022;11:555.
Chatterjee S, Matas AJ, Isaacson T, Kehlet C, Rose JKC, Stark RE. Solid-state13 C NMR delineates the architectural design of biopolymers in native and genetically altered tomato fruit cuticles. Biomacromol. 2016;17:215–24.
Pollard M, Beisson F, Li Y, Ohlrogge JB. Building lipid barriers: biosynthesis of cutin and suberin. Trends Plant Sci. 2008;13:236–46.
Cosgrove DJ, Dupree P, Gomez ED, Haigler CH, Kubicki JD, Zimmer J. How many glucan chains form plant cellulose microfibrils? A mini review. Biomacromol. 2024;25:6357–66.
Chen C-L, Robert D. Characterization of lignin by 1H and 13C NMR spectroscopy. In: Methods in enzymology. Academic Press; 1988. p. 137–74.
Gao X, Laskar DD, Zeng J, Helms GL, Chen S. A 13C CP/MAS-based nondegradative method for lignin content analysis. ACS Sustain Chem Eng. 2015;3:153–62.
Cipriano DF, Chinelatto LS, Nascimento SA, Rezende CA, de Menezes SMC, Freitas JCC. Potential and limitations of 13C CP/MAS NMR spectroscopy to determine the lignin content of lignocellulosic feedstock. Biomass Bioenerg. 2020;142: 105792.
Happs RM, Addison B, Doeppke C, Donohoe BS, Davis MF, Harman-Ware AE. Comparison of methodologies used to determine aromatic lignin unit ratios in lignocellulosic biomass. Biotechnol Biofuels. 2021;14:58.
Fu L, McCallum SA, Miao J, Hart C, Tudryn GJ, Zhang F, et al. Rapid and accurate determination of the lignin content of lignocellulosic biomass by solid-state NMR. Fuel. 2015;141:39–45.
Manders WF. Solid-state 13 C NMR determination of the syringyl/guaiacyl ratio in hardwoods. Holzforschung. 1987;41:13–8.
Nomura S, Kugo Y, Erata T. 13C NMR and XRD studies on the enhancement of cellulose II crystallinity with low concentration NaOH post-treatments. Cellulose. 2020;27:3553–63.
Yoo CG, Pu Y, Nelson R, Ragauskas AJ. Cellulose crystallinity index (CrI) of switchgrass measured by solid-state NMR. Oak Ridge National Laboratory (ORNL); 2024.
Foston MB, Hubbell CA, Ragauskas AJ. Cellulose isolation methodology for NMR analysis of cellulose ultrastructure. Materials (Basel). 2011;4:1985–2002.
Li M, Pu Y, Yoo CG, Gjersing E, Decker SR, Doeppke C, et al. Study of traits and recalcitrance reduction of field-grown COMT down-regulated switchgrass. Biotechnol Biofuels. 2017;10:12.
Yan L, Pu Y, Bowden M, Ragauskas AJ, Yang B. Physiochemical characterization of lignocellulosic biomass dissolution by flowthrough pretreatment. ACS Sustain Chem Eng. 2016;4:219–27.
Park S, Baker JO, Himmel ME, Parilla PA, Johnson DK. Cellulose crystallinity index: measurement techniques and their impact on interpreting cellulase performance. Biotechnol Biofuels. 2010;3:1–10.
Yoo CG, Meng X, Pu Y, Ragauskas AJ. The critical role of lignin in lignocellulosic biomass conversion and recent pretreatment strategies: a comprehensive review. Biores Technol. 2020;301: 122784.
Hossain MDM, Aldous L. Ionic liquids for lignin processing: dissolution, isolation, and conversion. Aust J Chem. 2012;65:1465–77.
Wu Z, Peng K, Zhang Y, Wang M, Yong C, Chen L, et al. Lignocellulose dissociation with biological pretreatment towards the biochemical platform: a review. Mater Today Bio. 2022;16: 100445.
Lima MA, Gomez LD, Steele-King CG, Simister R, Bernardinelli OD, Carvalho MA, et al. Evaluating the composition and processing potential of novel sources of Brazilian biomass for sustainable biorenewables production. Biotechnol Biofuels. 2014;7:1–19.
Sun Q, Foston M, Meng X, Sawada D, Pingali SV, O’Neill HM, et al. Effect of lignin content on changes occurring in poplar cellulose ultrastructure during dilute acid pretreatment. Biotechnol Biofuels. 2014;7:150.
Hossain MDM, Rawal A, Aldous L. Aprotic vs protic ionic liquids for lignocellulosic biomass pretreatment: anion effects, enzymatic hydrolysis, solid-state NMR, distillation, and recycle. ACS Sustain Chem Eng. 2019;7:11928–36.
Wi SG, Cho EJ, Lee D-S, Lee SJ, Lee YJ, Bae H-J. Lignocellulose conversion for biofuel: a new pretreatment greatly improves downstream biocatalytic hydrolysis of various lignocellulosic materials. Biotechnol Biofuels. 2015;8:1–11.
Sathitsuksanoh N, Zhu Z, Wi S, Percival Zhang Y-H. Cellulose solvent-based biomass pretreatment breaks highly ordered hydrogen bonds in cellulose fibers of switchgrass. Biotech Bioeng. 2011;108:521–9.
Rollin JA, Zhu Z, Sathitsuksanoh N, Zhang Y-HP. Increasing cellulose accessibility is more important than removing lignin: a comparison of cellulose solvent-based lignocellulose fractionation and soaking in aqueous ammonia. Biotech Bioeng. 2011;108:22–30.
Husson E, Buchoux S, Avondo C, Cailleu D, Djellab K, Gosselin I, et al. Enzymatic hydrolysis of ionic liquid-pretreated celluloses: contribution of CP-MAS 13C NMR and SEM. Biores Technol. 2011;102:7335–42.
Qu T, Zhang X, Gu X, Han L, Ji G, Chen X, et al. Ball milling for biomass fractionation and pretreatment with aqueous hydroxide solutions. ACS Sustain Chem Eng. 2017;5:7733–42.
Ekwe NB, Tyufekchiev MV, Salifu AA, Tompsett GA, LeClerc HO, Belden ER, et al. Mechanochemical pretreatment for waste-free conversion of bamboo to simple sugars: utilization of available resources for developing economies. Adv Sustain Syst. 2022;6:2100286.
Yuan S, Tyufekchiev MV, Timko MT, Schmidt-Rohr K. Direct quantification of the degree of polymerization of hydrolyzed cellulose by solid-state NMR spectroscopy. Cellulose. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s10570-022-04433-9.
Henuningson JA, Newman DSIRRH. A CP/MAS 13C NMR study of the effect of steam explosion processes on wood composition and structure. J Wood Chem Technol. 1985;5:159–88.
Yamashiki T, Matsui T, Saitoh M, Okajima K, Kamide K, Sawada T. Characterisation of cellulose treated by the steam explosion method. Part 1: Influence of cellulose resources on changes in morphology, degree of polymerisation, solubility and solid structure. Brit Poly J. 1990;22:73–83.
Nishida M, Tanaka T, Miki T, Ito T, Kanayama K. Multi-scale instrumental analyses for structural changes in steam-treated bamboo using a combination of several solid-state NMR methods. Ind Crops Prod. 2017;103:89–98.
Feng Y, Lei B, Liang Y, Zhong H, Yin X, Qu J, et al. Changes in the microstructure and components of Eulaliopsis binata treated by continuous screw extrusion steam explosion. BioResources. 2016;11:9455–66.
Hou X, Wang Z, Sun J, Li M, Wang S, Chen K, et al. A microwave-assisted aqueous ionic liquid pretreatment to enhance enzymatic hydrolysis of Eucalyptus and its mechanism. Biores Technol. 2019;272:99–104.
Wang Z, Hou X, Sun J, Li M, Chen Z, Gao Z. Comparison of ultrasound-assisted ionic liquid and alkaline pretreatment of Eucalyptus for enhancing enzymatic saccharification. Biores Technol. 2018;254:145–50.
Ben H, Hao N, Liu Q, Ragauskas AJ. Solid-state NMR Investigation of bio-chars produced from biomass components and whole biomasses. Bioenerg Res. 2017;10:1036–44.
Brar KK, Espirito Santo MC, Pellegrini VOA, deAzevedo ER, Guimaraes FEC, Polikarpov I, et al. Enhanced hydrolysis of hydrothermally and autohydrolytically treated sugarcane bagasse and understanding the structural changes leading to improved saccharification. Biomass Bioenerg. 2020;139: 105639.
Chandel AK, Antunes FA, Silva MB, Da Silva S. Unraveling the structure of sugarcane bagasse after soaking in concentrated aqueous ammonia (SCAA) and ethanol production by Scheffersomyces (Pichia) stipitis. Biotechnol Biofuels. 2013;6:102.
Freitas JCC, Bonagamba TJ, Emmerich FG. Investigation of biomass- and polymer-based carbon materials using 13C high-resolution solid-state NMR. Carbon. 2001;39:535–45.
Sluiter JB, Michel KP, Addison B, Zeng Y, Michener W, Paterson AL, et al. Direct determination of cellulosic glucan content in starch-containing samples. Cellulose. 2021;28:1989–2002.
Bonanomi G, Incerti G, Giannino F, Mingo A, Lanzotti V, Mazzoleni S. Litter quality assessed by solid state 13C NMR spectroscopy predicts decay rate better than C/N and Lignin/N ratios. Soil Biol Biochem. 2013;56:40–8.
Martínez AT, González AE, Valmaseda M, Dale BE, Lambregts MJ, Haw JF. Solid-state NMR studies of lignin and plant polysaccharide degradation by fungi. Holzforschung. 1991;45:49–54.
Davis MF, Schroeder HA, Maciel GE. Solid-state 13C nuclear magnetic resonance studies of wood decay. III. Decay of colorado blue spruce and paper birch by Postia placenta. HFSG. 1994;48:301–7.
Melillo JM, Aber JD, Muratore JF. Nitrogen and lignin control of hardwood leaf litter decomposition dynamics. Ecology. 1982;63:621–6.
Taylor BR, Parkinson D, Parsons WFJ. Nitrogen and lignin content as predictors of litter decay rates: a microcosm test. Ecology. 1989;70:97–104.
D’Aizat N, Saad NM, Ahmad Usuldin SR, Vayabari DAG, Ilham Z, Ibrahim MF, et al. Bioactivities of kenaf biomass extracts: a review. Processes. 2023;11:1178.
Schmidt-Rohr K, Spiess HW. Multidimensional solid-state NMR and polymers. Academic Press; 1994.
Walker TW, Kuch N, Vander Meulen KA, Clewett CFM, Huber GW, Fox BG, et al. Solid-State NMR studies of solvent-mediated, acid-catalyzed woody biomass pretreatment for enzymatic conversion of residual cellulose. ACS Sustainable Chem Eng. 2020;8:6551–63.
Ghosh M, Kango N, Dey KK. Investigation of the internal structure and dynamics of cellulose by 13C-NMR relaxometry and 2DPASS-MAS-NMR measurements. J Biomol NMR. 2019;73:601–16.
Sparrman T, Svenningsson L, Sahlin-Sjövold K, Nordstierna L, Westman G, Bernin D. A revised solid-state NMR method to assess the crystallinity of cellulose. Cellulose. 2019;26:8993–9003.
Zumbulyadis N. Selective carbon excitation and the detection of spatial heterogeneity in cross-polarization magic-angle-spinning NMR. J Magn Reson. 1969;1983(53):486–94.
Newman RH, Condron LM. Separating subspectra from cross-polarization magic-angle spinning nuclear magnetic resonance spectra by proton spin relaxation editing. Solid State Nucl Magn Reson. 1995;4:259–66.
Preston CM, Smernik RJ, Powers RF, McColl JG, McBeath TM. The decomposition of windrowed, chipped logging slash and tree seedling response: a plant growth and nuclear magnetic resonance spectroscopy study. Org Geochem. 2011;42:936–46.
Chen Q, Schmidt-Rohr K. Measurement of the local 1H spin-diffusion coefficient in polymers. Solid State Nucl Magn Reson. 2006;29:142–52.
Ibbett RN, Domvoglou D, Fasching M. Characterisation of the supramolecular structure of chemically and physically modified regenerated cellulosic fibres by means of high-resolution Carbon-13 solid-state NMR. Polymer. 2007;48:1287–96.
Scheler U. Solid polymers. In: Solid-state NMR spectroscopy principles and applications. John Wiley & Sons, Ltd; 2001. p. 483–511. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/9780470999394.ch10.
Newman RH. Estimation of the lateral dimensions of cellulose crystallites using 13C NMR signal strengths. Solid State Nucl Magn Reson. 1999;15:21–9.
Fan LT, Lee Y-H, Beardmore DH. Mechanism of the enzymatic hydrolysis of cellulose: Effects of major structural features of cellulose on enzymatic hydrolysis. Biotechnol Bioeng. 1980;22:177–99.
Hall M, Bansal P, Lee JH, Realff MJ, Bommarius AS. Cellulose crystallinity—a key predictor of the enzymatic hydrolysis rate. FEBS J. 2010;277:1571–82.
Peciulyte A, Karlström K, Larsson PT, Olsson L. Impact of the supramolecular structure of cellulose on the efficiency of enzymatic hydrolysis. Biotechnol Biofuels. 2015;8:56.
Meng X, Pu Y, Yoo CG, Li M, Bali G, Park D-Y, et al. An in-depth understanding of biomass recalcitrance using natural poplar variants as the feedstock. Chemsuschem. 2017;10:139–50.
Tyufekchiev M, Kolodziejczak A, Duan P, Foston M, Schmidt-Rohr K, Timko TM. Reaction engineering implications of cellulose crystallinity and water-promoted recrystallization. Green Chem. 2019;21:5541–55.
Gamble GR, Sethuraman A, Akin DE, Eriksson KE. Biodegradation of lignocellulose in Bermuda grass by white rot fungi analyzed by solid-state 13C nuclear magnetic resonance. Appl Environ Microbiol. 1994;60:3138–44.
Mao J-D, Schmidt-Rohr K. Recoupled long-range C-H dipolar dephasing in solid-state NMR, and its use for spectral selection of fused aromatic rings. J Magn Reson. 2003;162:217–27.
Mao J, Holtman KM, Scott JT, Kadla JF, Schmidt-Rohr K. Differences between lignin in unprocessed wood, milled wood, mutant wood, and extracted lignin detected by 13C solid-state NMR. J Agric Food Chem. 2006;54:9677–86.
Rezende CA, de Lima MA, Maziero P, deAzevedo ER, Garcia W, Polikarpov I. Chemical and morphological characterization of sugarcane bagasse submitted to a delignification process for enhanced enzymatic digestibility. Biotechnol Biofuels. 2011;4:54.
Opella SJ, Frey MH, Cross TA. Detection of individual carbon resonances in solid proteins. J Am Chem Soc. 1979;101:5856–7.
Mao J-D, Holtman KM, Franqui-Villanueva D. Chemical structures of corn stover and its residue after dilute acid prehydrolysis and enzymatic hydrolysis: insight into factors limiting enzymatic hydrolysis. J Agric Food Chem. 2010;58:11680–7.
Clauss J, Schmidt-Rohr K, Spiess HW. Determination of domain sizes in heterogeneous polymers by solid-state NMR. Acta Polymer. 1993;44:1–17.
Ahvazi BC, Argyropoulos DS. Proton spin–lattice relaxation time measurements of solid wood and its constituents as a function of pH: Part II. Solid State Nucl Magn Reson. 1999;15:49–57.
Nishida M, Tanaka T, Miki T, Hayakawa Y, Kanayama K. Instrumental analyses of nanostructures and interactions with water molecules of biomass constituents of Japanese cypress. Cellulose. 2017;24:5295–312.
Argyropoulos DS, Morin FG. Probing the macromolecular structure of wood and pulps with proton spin-lattice relaxation time measurements in the solid state. Wood SciTechnol. 1995;29:19–30.
Garvey CJ, Simon GP, Whittaker AK, Parker IH. Moisture-activated dynamics on crystallite surfaces in cellulose. Colloid Polym Sci. 2019;297:521–7.
Hediger S, Emsley L, Fischer M. Solid-state NMR characterization of hydration effects on polymer mobility in onion cell-wall material. Carbohyd Res. 1999;322:102–12.
Santoni I, Callone E, Sandak A, Sandak J, Dirè S. Solid state NMR and IR characterization of wood polymer structure in relation to tree provenance. Carbohyd Polym. 2015;117:710–21.
Newman RH. Nuclear magnetic resonance study of spatial relationships between chemical components in wood cell walls. Holzforschung. 1992;46:205–10.
Leroy A, Falourd X, Foucat L, Méchin V, Guillon F, Paës G. Evaluating polymer interplay after hot water pretreatment to investigate maize stem internode recalcitrance. Biotechnol Biofuels. 2021;14:164.
Dick-Pérez M, Zhang Y, Hayes J, Salazar A, Zabotina OA, Hong M. Structure and interactions of plant cell-wall polysaccharides by two- and three-dimensional magic-angle-spinning solid-state NMR. Biochemistry. 2011;50:989–1000.
Dick-Perez M, Wang T, Salazar A, Zabotina OA, Hong M. Multidimensional solid-state NMR studies of the structure and dynamics of pectic polysaccharides in uniformly 13C-labeled Arabidopsis primary cell walls: Pectin structure in plant cell walls from solid-state NMR. Magn Reson Chem. 2012;50:539–50.
Torchia DA. The measurement of proton-enhanced carbon-13 T1 values by a method which suppresses artifacts. J Magn Reson. 1969;1978(30):613–6.
Ghosh M, Prajapati BP, Suryawanshi RK, Kishor Dey K, Kango N. Study of the effect of enzymatic deconstruction on natural cellulose by NMR measurements. Chem Phys Lett. 2019;727:105–15.
Martin AF, Tobimatsu Y, Kusumi R, Matsumoto N, Miyamoto T, Lam PY, et al. Altered lignocellulose chemical structure and molecular assembly in cinnamyl alcohol dehydrogenase-deficient rice. Sci Rep. 2019;9:17153.
Martin AF, Tobimatsu Y, Lam PY, Matsumoto N, Tanaka T, Suzuki S, et al. Lignocellulose molecular assembly and deconstruction properties of lignin-altered rice mutants. Plant Physiol. 2023;191:70–86.
Kohn B, Davis M, Maciel GE. In situ study of dilute H2SO4 pretreatment of 13 C-enriched poplar wood, using 13C NMR. Energy Fuels. 2011;25:2301–13.
Kobayashi T, Kohn B, Holmes L, Faulkner R, Davis M, Maciel GE. Molecular-level consequences of biomass pretreatment by dilute sulfuric acid at various temperatures. Energy Fuels. 2011;25:1790–7.
Nishida M, Tanaka T, Miki T, Shigematsu I, Kanayama K, Kanematsu W. Study of nanoscale structural changes in isolated bamboo constituents using multiscale instrumental analyses. J Appl Polym Sci. 2014. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/app.40243.
Zuckerstätter G, Terinte N, Sixta H, Schuster KC. Novel insight into cellulose supramolecular structure through 13C CP-MAS NMR spectroscopy and paramagnetic relaxation enhancement. Carbohyd Polym. 2013;93:122–8.
Novotny EH, Garcia RHS, deAzevedo ER. Pulse sequence induced variability combined with multivariate analysis as a potential tool for 13C solid-state NMR signals separation, quantification, and classification. J Magn Resonan Open. 2023;14–15: 100089.
White PB, Wang T, Park YB, Cosgrove DJ, Hong M. Water-polysaccharide interactions in the primary cell wall of Arabidopsis thaliana from polarization transfer solid-state NMR. J Am Chem Soc. 2014;136:10399–409.
Chakraborty A, Fernando LD, Fang W, Dickwella Widanage MC, Wei P, Jin C, et al. A molecular vision of fungal cell wall organization by functional genomics and solid-state NMR. Nat Commun. 2021;12:6346.
Dickwella Widanage MC, Gautam I, Sarkar D, Mentink-Vigier F, Vermaas JV, Ding S-Y, et al. Adaptative survival of Aspergillus fumigatus to echinocandins arises from cell wall remodeling beyond β−1,3-glucan synthesis inhibition. Nat Commun. 2024;15:6382.
Huster D, Yao X, Hong M. Membrane protein topology probed by1 h spin diffusion from lipids using solid-state NMR spectroscopy. J Am Chem Soc. 2002;124:874–83.
Kolodziejski W, Klinowski J. Kinetics of cross-polarization in solid-state NMR: a guide for chemists. Chem Rev. 2002;102:613–28.
Hong M, Gross JD, Rienstra CM, Griffin RG, Kumashiro KK, Schmidt-Rohr K. Coupling amplification in 2D MAS NMR and its application to torsion angle determination in peptides. J Magn Reson. 1997;129:85–92.
Hester RK, Ackerman JL, Neff BL, Waugh JS. Separated local field spectra in NMR: determination of structure of solids. Phys Rev Lett. 1976;36:1081–3.
deAzevedo ER, Saalwachter K, Pascui O, de Souza AA, Bonagamba TJ, Reichert D. Intermediate motions as studied by solid-state separated local field NMR experiments. J Chem Phys. 2008;128: 104505.
Bielecki A, Kolbert AC, Levitt MH. Frequency-switched pulse sequences: Homonuclear decoupling and dilute spin NMR in solids. Chem Phys Lett. 1989;155:341–6.
Mansfield P, Orchard MJ, Stalker DC, Richards KHB. Symmetrized multipulse nuclear-magnetic-resonance experiments in solids: measurement of the chemical-shift shielding tensor in some compounds. Phys Rev B. 1973;7:90–105.
Wang L-Q, Shin Y, Samuels WD, Exarhos GJ, Moudrakovski IL, Terskikh VV, et al. Magnetic resonance studies of hierarchically ordered replicas of wood cellular structures prepared by surfactant-mediated mineralization. J Phys Chem B. 2003;107:13793–802.
Capitani D, Proietti N, Ziarelli F, Segre AL. NMR study of water-filled pores in one of the most widely used polymeric material: the paper. Macromolecules. 2002;35:5536–43.
Spinella A, Chillura Martino DF, Saladino ML, Sammartino F, Caruso F, Caponetti E. Solid state NMR investigation of the roman Acqualadroni rostrum: tenth year assessment of the consolidation treatment of the wooden part. Cellulose. 2021;28:1025–38.
Le Brech Y, Raya J, Delmotte L, Brosse N, Gadiou R, Dufour A. Characterization of biomass char formation investigated by advanced solid state NMR. Carbon. 2016;108:165–77.
Schmidt-Rohr K, Clauss J, Spiess HW. Correlation of structure, mobility, and morphological information in heterogeneous polymer materials by two-dimensional wideline-separation NMR spectroscopy. Macromolecules. 1992;25:3273–7.
Yan B, Stark RE. A WISE NMR approach to heterogeneous biopolymer mixtures: dynamics and domains in wounded potato tissues. Macromolecules. 1998;31:2600–5.
Yu B, Vengadesan G, Wang H, Jashi L, Yefremov T, Tian S, et al. Magic-angle spinning nmr studies of cell wall bound aromatic−aliphatic biopolyesters associated with strengthening of intercellular adhesion in potato (Solanum tuberosum L.) tuber parenchyma. Biomacromol. 2006;7:937–44.
Round AN, Yan B, Dang S, Estephan R, Stark RE, Batteas JD. The influence of water on the nanomechanical behavior of the plant biopolyester cutin as studied by AFM and solid-state NMR. Biophys J. 2000;79:2761–7.
Bielecki A, Kolbert AC, De Groot HJM, Griffin RG, Levitt MH. Frequency-switched Lee-Goldburg sequences in solids. In: Warren WS, editor. Advances in magnetic and optical resonance. Academic Press; 1990. p. 111–24.
van Rossum B-J, Förster H, de Groot HJM. High-field and high-speed CP-MAS13C NMR heteronuclear dipolar-correlation spectroscopy of solids with frequency-switched Lee-Goldburg homonuclear decoupling. J Magn Reson. 1997;124:516–9.
Holtman KM, Chen N, Chappell MA, Kadla JF, Xu L, Mao J. Chemical structure and heterogeneity differences of two lignins from loblolly pine as investigated by advanced solid-state NMR spectroscopy. J Agric Food Chem. 2010;58:9882–92.
Le Brech Y, Delmotte L, Raya J, Brosse N, Gadiou R, Dufour A. High resolution solid state 2D NMR analysis of biomass and biochar. Anal Chem. 2015;87:843–7.
Knicker H, Velasco-Molina M, Knicker M. 2D solid-state HETCOR 1H–13C NMR experiments with variable cross polarization times as a tool for a better understanding of the chemistry of cellulose-based pyrochars—a tutorial. Appl Sci. 2021;11:8569.
Nishiyama Y. Fast magic-angle sample spinning solid-state NMR at 60–100 kHz for natural abundance samples. Solid State Nucl Magn Reson. 2016;78:24–36.
Yuan ECY, Huang SJ, Huang HC, Sinkkonen J, Oss A, Org ML, et al. Faster magic angle spinning reveals cellulose conformations in woods. Chem Commun. 2021;57:4110–3.
Mentink-Vigier F, Marin-Montesinos I, Jagtap AP, Halbritter T, Van Tol J, Hediger S, et al. Computationally assisted design of polarizing agents for dynamic nuclear polarization enhanced NMR: the AsymPol family. J Am Chem Soc. 2018;140:11013–9.
Scott FJ, Eddy S, Gullion T, Mentink-Vigier F. Sorbitol-based glass matrices enable dynamic nuclear polarization beyond 200 K. J Phys Chem Lett. 2024;15:8743–51.
Acknowledgements
The authors thank Besiki Kazaishvili of NREL's Communications Team for assistance with Figure 1 illustration. The authors also thank Renee Happs and Mark Davis for insightful discussions and historical direction of NREL's NMR Facility.
Funding
This work was authored in part by the National Renewable Energy Laboratory, operated by Alliance for Sustainable Energy, LLC, for the U.S. Department of Energy (DOE) under Contract No. DE-AC36-08GO28308. This material is based upon work supported by the Center for Bioenergy Innovation (CBI), U.S. Department of Energy, Office of Science, Biological and Environmental Research Program under Award Number ERKP886. The views expressed in the article do not necessarily represent the views of the DOE or the U.S. Government. The U.S. Government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes.
Author information
Authors and Affiliations
Contributions
BA conceptualized the manuscript. BA and MCD wrote and edited the manuscript, prepared all figures, and provided data. YP provided specific sections, edited the full manuscript, and provided data. AEW and AR provided crucial review and contributed to the editing of all sections. All authors read and approved the final manuscript.
Corresponding author
Ethics declarations
Ethics approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Additional information
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Addison, B., Dickwella Widange, M.C., Pu, Y. et al. Solid-state NMR at natural isotopic abundance for bioenergy applications. Biotechnol. Biofuels Bioprod. 18, 46 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13068-025-02648-z
Received:
Accepted:
Published:
DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13068-025-02648-z