Fact-checked by Grok 2 weeks ago

Lactide

Lactide is a cyclic diester and dimer of , with the molecular formula C₆H₈O₄ and systematic name 3,6-dimethyl-1,4-dioxane-2,5-dione, primarily serving as a for the of poly() (PLA), a . First synthesized by Théophile-Jules Pelouze in 1845 through dehydration of , it has since become central to production. Due to the chirality of , lactide exists in several stereoisomers: L-lactide and D-lactide (enantiomers with melting points of 95–100°C and specific rotations of −260° to −300° and +260° to +300°, respectively), meso-lactide ( 53–54°C), and rac-lactide (a 1:1 mixture of L- and D-lactide with around 125°C and specific rotation near 0°). These isomers appear as white crystalline solids and exhibit and bioabsorbability, making L-lactide particularly suitable for biomedical applications. Lactide is typically produced via a process involving the polycondensation of to form low-molecular-weight oligomers, followed by their catalytic at temperatures above 200°C under reduced , using catalysts such as tin(II) octoate to promote back-biting reactions and minimize . Alternative direct synthesis routes from have been explored but remain less common due to challenges in selectivity and . The compound's key applications leverage PLA's properties, including use in resorbable medical devices like prostheses, surgical membranes, and systems, as well as in when blended with other polymers. Its production and enable the creation of high-molar-mass PLA with tunable degradation rates, supporting eco-friendly alternatives to petroleum-based plastics.

Introduction

Definition and Overview

Lactide is a cyclic diester derived from two molecules of (2-hydroxypropanoic acid) through esterification and processes. This results in a dilactide structure characterized by a six-membered ring incorporating two ester groups and methyl substituents. Lactide serves as a key intermediate in the production of high-molecular-weight polymers, particularly (), via . As a renewable, bio-based chemical, lactide is obtained from produced by the of carbohydrates such as or , offering a sustainable alternative to petroleum-derived polymers in applications. This bio-origin contributes to its significance in developing biodegradable plastics that reduce reliance on fossil fuels. Lactide exists in various stereoisomeric forms, which influence the properties of the resulting polymers.

Historical Development

The synthesis of lactide, the cyclic diester of , was first described in 1845 by French chemist Théophile-Jules Pelouze, who obtained it through the of lactic acid under reduced pressure, initially as a byproduct in explorations. This early discovery positioned lactide as a curiosity, with limited practical application until advancements in emerged decades later. In the 1930s, interest in lactide grew with ' research at , where he and collaborators demonstrated its to form (PLA) in 1932, marking the initial attempts to produce polyesters from renewable sources. However, these efforts yielded only low-molecular-weight polymers due to impurities in the lactide , constraining commercial viability. Progress accelerated in the 1950s when developed enhanced purification methods for lactide in 1954, allowing for the synthesis of high-molecular-weight PLA and establishing the lactide route as a key pathway for polyester production. The saw a surge in industrial focus, driven by demands, as and Dow Chemical formed a in 1997 to scale lactide-based production from corn-derived , culminating in NatureWorks' commercial facility opening in Blair, , in 2001 with an initial capacity of 140,000 tonnes per year. This milestone shifted lactide from laboratory novelty to a viable precursor, emphasizing renewable feedstocks. Post-2010 developments have emphasized , particularly in chemical where end-of-life is depolymerized back to lactide using efficient catalysts like complexes, achieving near-quantitative yields under mild conditions to support circular processes. These advances, including solvent-free and enzymatic-assisted methods, have reduced energy inputs and impurities, enhancing the environmental profile of lactide-based materials. More recently, as of 2025, industrial production has expanded with NatureWorks advancing construction of a new facility in , , announced in 2023, aiming to increase global capacity and support bio-based polymer demand. Additionally, new projects in , such as those focusing on lactide and production, were reported in 2024, further scaling sustainable .

Chemical Structure and Properties

Molecular Formula and Bonding

Lactide possesses the molecular formula C₆H₈O₄ and a of 144.12 g/mol. The features a six-membered 1,4-dioxane-2,5-dione ring, with two chiral carbon atoms at positions 3 and 6, each substituted with a , and linked by two functionalities comprising C=O and C-O bonds. In the solid state, the exhibits a flattened conformation, where the sp² hybridization of the carbonyl carbon atoms promotes partial planarity among the heavy atoms while maintaining overall rigidity. The polar groups generate a significant , oriented along the molecular axis due to the electron-withdrawing nature of the carbonyl oxygens. Intramolecular is precluded by the absence of hydroxyl groups, though intermolecular C-H···O interactions occur between methylene hydrogens and carbonyl oxygens in the crystal lattice, with H···O distances of 2.3–2.4 Å. Isotopically labeled variants, such as ¹³C-enriched lactide, facilitate detailed NMR investigations, enabling tracking of carbon positions in mechanistic studies of ring-opening reactions and formation.

Physical and Thermal Properties

Lactide is typically observed as a white crystalline solid. The of lactide varies depending on its stereoisomeric form; (R,R)- and (S,S)-lactide exhibit melting points in the range of 95–100 °C, while meso-lactide has a significantly lower of 53–54 °C. The of L- and D-lactide is approximately 255 °C at 760 mmHg, though it decomposes prior to reaching this temperature under standard conditions. The density of (R,R)- and (S,S)-lactide is reported in the range of 1.32–1.38 g/cm³ at 20 °C. Lactide demonstrates good solubility in polar organic solvents such as and , moderate solubility in acetone and , solubility in where it undergoes to , and insolubility in non-polar solvents like . Lactide exhibits thermal stability up to approximately 200 °C, with complete observed around 233 °C as determined by ; as a crystalline material, it lacks a distinct temperature. In , lactide shows characteristic absorption bands including a strong carbonyl (C=O) stretch at approximately 1754 cm⁻¹, along with peaks at 935, 1056, 1093, 1240, and 1266 cm⁻¹. ¹H NMR of lactide reveals a for the methyl protons at δ 1.6 and a quadruplet for the methine protons at δ 5.0 .

Stereoisomers

Lactide exists in three stereoisomeric forms due to the of its precursor , which possesses a stereogenic center at the alpha carbon. These isomers are (3R,6R)-3,6-dimethyl-1,4-dioxane-2,5-dione, commonly known as (R,R)-lactide or D-lactide, which is dextrorotatory; (3S,6S)-3,6-dimethyl-1,4-dioxane-2,5-dione, or (S,S)-lactide (L-lactide), which is levorotatory; and the meso form, (3R,6S)-3,6-dimethyl-1,4-dioxane-2,5-dione (or equivalently (3S,6R)), which is achiral due to an internal plane of symmetry. A , rac-lactide (1:1 L- and D-lactide), has a of 122–128 °C and near 0°.(https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9032396/)[](https://pubs.rsc.org/en/content/articlehtml/2015/py/c4py01572j) The configurations of these stereoisomers are directly derived from the enantiomerically pure s used in their synthesis: (R,R)-lactide forms from two molecules of (R)-, (S,S)-lactide from two (S)- molecules, and meso-lactide from a combination of one (R)- and one (S)- unit. As enantiomers, (R,R)- and (S,S)-lactide exhibit identical physical properties, including melting points of 95–98 °C and solubilities in organic solvents such as and ; in contrast, meso-lactide has a lower melting point of 52–54 °C and demonstrates higher and faster rates in aqueous environments due to its reduced crystallinity. The optical activity of these isomers is a key distinguishing feature: the [α]_D for (R,R)-lactide is approximately +287° to +300° (in at 25 °C), for (S,S)-lactide it is -287° to -300°, and meso-lactide shows no ([α]_D = 0°) owing to its achirality. Epimerization of lactide stereoisomers can occur through base-catalyzed mechanisms, involving enolization or cleavage, which facilitates interconversion between the enantiomers and the meso form, particularly under elevated temperatures (e.g., above 200 °C) or in the presence of catalysts like metal alkoxides. This process reduces optical purity and is a challenge in maintaining during handling or synthesis. In production, (S,S)-lactide predominates when derived from L-lactic acid, which is primarily obtained via microbial fermentation of carbohydrates using strains like Lactobacillus that favor the (S)-enantiomer; industrial processes often yield mixtures containing varying proportions of all three isomers due to partial racemization during depolymerization steps. These stereoisomers influence the tacticity and crystallinity of the resulting polylactic acid (PLA) polymers, with pure (S,S)-lactide yielding highly isotactic, semicrystalline PLA.

Synthesis

Preparation from Lactic Acid

Lactide is synthesized from through a two-step process involving the condensation of two molecules into a cyclic dimer, accompanied by the elimination of two molecules. The basic can be represented as: $2 \ \ce{CH3CH(OH)COOH} \rightarrow \ce{[CH3CHCOOCHCOOCHCH3]} + 2 \ \ce{H2O} This simplified equation depicts the formation of the six-membered ring of lactide. The first step entails polycondensation of lactic acid to form low-molecular-weight oligo(lactic acid) with number-average molecular weights (Mₙ) typically ranging from 800 to 3000 Da. This reaction is conducted using acid catalysts such as tin(II) chloride (SnCl₂) or tin(II) octoate at temperatures ranging from 130–180 °C under vacuum to facilitate water removal and drive the equilibrium toward oligomer formation. For instance, the process performed gradually—starting at 120 °C for 1 hour, then 150 °C for 2 hours, and 180 °C for 2 hours—yields oligomers with number-average molecular weights (Mₙ) of 2272–2390 Da. Similarly, tin octoate (0.25–1.00 wt%) at 140–200 °C and 510–10 mbar pressure produces oligomers with Mₙ up to 1995 Da after 3 hours under nitrogen atmosphere. In the second step, the oligo(lactic acid) undergoes cracking or to generate lactide monomer. This is achieved via thermal or catalytic methods at 200–250 °C, often under reduced pressure to promote volatilization of the product. Catalysts such as zinc oxide (ZnO) or tin octoate are employed; for example, with 0.1 wt% SnCl₂ at 210 °C and 76 for 3 hours up to 78.8% crude lactide, while ZnO nanoparticles at 220 °C and 3 kPa produce 77–80% . Tin octoate-mediated at 210 °C and 5–10 mbar achieves 67–69% raw lactide . Laboratory-scale generally range from 50–70%, with side products including linear oligomers and residual that can complicate separation. Purification of the crude lactide is typically accomplished by to isolate the high-purity . At approximately 10 mmHg, lactide distills at around 140 °C, allowing separation from higher-boiling impurities like oligomers while minimizing . This process yields polymer-grade lactide with purity exceeding 99%, essential for subsequent applications.

Industrial Methods and Purification

The primary industrial route for lactide production begins with the of glucose or other carbohydrates using bacterial strains such as to yield L- with an enantiomeric excess typically exceeding 95%. This is then converted to lactide through a two-step process: first, catalytic prepolymerization forms low-molecular-weight oligomers under vacuum at elevated temperatures (around 150–200°C), followed by in the presence of catalysts like tin(II) octoate to crack the oligomers into cyclic lactide monomers. The NatureWorks process, operational since 2002, exemplifies this approach, employing a continuous, proprietary method that integrates , purification, and lactide formation for efficient scalability. Advanced methods focus on enhancing selectivity and reducing energy demands, including microwave-assisted depolymerization of lactic acid oligomers, which accelerates the reaction under solvent-free conditions and achieves yields up to 80% in shorter times compared to conventional heating. Solvent-free processes using heterogeneous catalysts, such as Sn-beta zeolites or SnO₂/SiO₂ nanocomposites, enable one-step continuous synthesis at moderate temperatures (110–240°C) and , improving catalyst recyclability and minimizing side products. These innovations, often explored in pilot scales, aim for higher L-lactide purity while addressing limitations of . Purification and yield optimization are critical for commercial viability, with continuous separating lactide (boiling point ~140 °C under reduced pressure) to achieve 99% purity, while recycling unreacted and water via esterification loops boosts overall yields to over 90%. Energy-efficient processes, including , reduce operational costs to approximately $1.50–2.00 per kg by lowering energy input and enabling impurity removal without excessive solvent use. Key challenges include minimizing meso-lactide formation (targeting <5% to preserve stereoregularity in downstream polymers) and eliminating impurities like lactoyllactic acid through combined and crystallization steps. Global lactide production capacity reached approximately 670,000 tons per year as of 2021, with projections exceeding 1 million tons by 2025, driven by demand for polylactic acid, with major facilities in the USA (e.g., NatureWorks/Corbion), China, and Europe (e.g., TotalEnergies). Patents from Corbion and TotalEnergies underscore ongoing innovations in process integration and catalyst design for cost-effective scaling.

Polymerization

Ring-Opening Polymerization Mechanism

The ring-opening polymerization (ROP) of proceeds primarily through a coordination-insertion mechanism, in which the lactide monomer coordinates to the active site of the initiator via its ester oxygen, positioning the carbonyl group for nucleophilic attack by an alkoxide chain end. This attack forms a tetrahedral intermediate, leading to cleavage of the acyl-oxygen bond and ring opening, driven by the ring strain inherent in lactide's six-membered 1,4-dioxane-2,5-dione structure. The opened lactide unit inserts into the metal-oxygen (or equivalent) bond, regenerating the active alkoxide species for further monomer addition. Propagation occurs through successive additions of lactide monomers to the active chain end, typically an alkoxide or ester terminus, extending the polyester chain. Transesterification reactions may also take place, where the chain end attacks internal ester linkages, potentially redistributing monomer units along the polymer backbone. The overall reaction can be represented as: n \ce{(CH3CHOCO)2O} + \ce{ROH} \rightarrow \ce{HO-[(CH3CHCO)O]_n-OR} where \ce{(CH3CHOCO)2O} denotes , \ce{ROH} is an alcohol initiator, and no byproducts form under ideal conditions. This process exemplifies living polymerization, characterized by linear kinetics where monomer consumption follows a first-order dependence on lactide concentration. Molecular weight is precisely controlled by the initiator-to-monomer ratio, yielding polymers with number-average molecular weights proportional to [ \text{lactide} ] / [ \text{initiator} ] \times \text{molecular weight of lactide}, and polydispersity indices (PDI) typically below 1.5 when employing effective initiators. Stereochemistry in the resulting polylactic acid (PLA) is dictated by the lactide stereoisomer employed. Enantiopure (R,R)- or (S,S)-lactide yields highly isotactic PLA with regular stereocenters, exhibiting high crystallinity and melting points up to 180°C. Racemic mixtures of (R,R)- and (S,S)-lactide produce atactic PLA with random stereosequences, resulting in amorphous polymers with lower glass transition temperatures around 55–60°C. Meso-lactide, comprising (R,S)-isomers, leads to syndiotactic PLA featuring alternating stereocenters, which can enhance mechanical properties compared to atactic forms. Stereocontrol is maintained through selective ring-opening at specific acyl-oxygen bonds, influenced by the initiator's chirality. Side reactions can compromise chain integrity and stereoregularity. Backbiting, where the active chain end intramolecularly attacks its own ester groups, generates cyclic oligomers and reduces linear polymer yield, particularly in dilute solutions or at elevated temperatures. Racemization occurs via enolization or epimerization at the α-carbon, especially under basic conditions or high temperatures (>150°C), leading to loss of optical purity and broader distributions in the product. These reactions are minimized in controlled living conditions but remain challenges in scaling .

Catalysts and Process Conditions

The ring-opening polymerization (ROP) of lactide to produce polylactic acid (PLA) predominantly employs metal-based catalysts, with stannous octoate (Sn(Oct)₂) being the most widely used due to its high activity, FDA approval, and ability to yield polymers with molar masses exceeding 100 kDa. Typically applied at concentrations of 0.01–0.1 mol%, Sn(Oct)₂ facilitates efficient polymerization under industrially relevant conditions, though higher loadings can lead to transesterification and reduced molecular weight control. Alternative catalysts include aluminum alkoxides, which achieve molar masses below 100 kDa but offer cost advantages, and zinc-based systems like ZnO or zinc 2-ethylhexanoate, noted for biocompatibility in biomedical applications. Rare-earth metal complexes, such as lanthanum-based initiators, provide enhanced biocompatibility and stereocontrol, enabling the production of isotactic PLA with minimal toxicity residues. Co-initiators are essential for controlled initiation in these systems, with alcohols such as commonly paired with Sn(Oct)₂ to promote chain growth and achieve molar masses over 1000 kDa, while can serve as a simple alternative but risks side reactions. For greener processes, enzymatic catalysts like s (e.g., Candida antarctica lipase B) enable mild ROP in organic solvents or bulk, offering high specificity and reduced metal contamination, though with slower rates compared to metal catalysts. Polymerization conditions typically involve or processes at temperatures of 130–180 °C for 1–24 hours, with vacuum applied to remove volatiles like unreacted and prevent . These parameters balance and polymer quality, as lower temperatures (around 120 °C) favor stereoregularity but extend times, while higher ones (up to 200 °C) accelerate conversion at the cost of potential chain scission. Optimizations focus on efficiency and , including heating, which accelerates rates by up to 10-fold compared to conventional methods through rapid, uniform energy transfer, enabling complete conversion in minutes at 140–180 °C. Solvent-free enhances environmental sustainability by eliminating organic solvent use and waste, while strict humidity control—via drying lactide to <0.01% water content—is critical to minimize and maintain high molecular weights. Recent advances as of 2023–2025 include low-toxicity catalysts such as (IV) salicylate complexes for efficient ROP of D,L-lactide and water-initiated systems using non-toxic salts, promoting greener and more accessible methods. Additionally, organocatalytic approaches have improved in meso-lactide . In commercial production, melt-phase ROP occurs continuously in twin-screw extruders, as exemplified by NatureWorks' Ingeo process, which operates at 180–210 °C under reduced pressure with Sn(Oct)₂ concentrations of 0.01–0.1 wt%, achieving circa 95% monomer conversion and PLA molar masses of 100,000–200,000 g/mol for high-performance applications.

Applications

In Polylactic Acid Production

Lactide serves as the primary for the industrial production of (PLA) via (ROP), a process that dominates commercial synthesis due to its efficiency in generating high-molecular-weight polymers. This route consumes the vast majority of produced lactide, enabling PLA with molecular weights exceeding 100,000 g/mol, far surpassing the limitations of direct condensation from , which typically yields polymers below 50,000 g/mol owing to challenges in water removal and chain extension. The ROP method also offers advantages in achieving higher purity, precise control over , and minimization of byproducts compared to direct polycondensation, which often results in oligomeric impurities and inconsistent chain lengths. PLA variants derived from lactide stereoisomers exhibit distinct material properties tailored for specific uses. Polymerization of L-lactide produces semi-crystalline poly(L-lactic acid) (PLLA), valued for its mechanical strength and thermal stability, while D-lactide yields poly(D-lactic acid) (PDLA), which can blend with PLLA to form stereocomplexes with elevated melting points (up to 230°C) and improved barrier properties. In contrast, racemic (D,L)-lactide results in amorphous poly(DL-lactic acid) (PDLLA), offering flexibility and faster degradation rates suitable for certain packaging applications. These variants allow customization of PLA's crystallinity and processability without altering the core ROP mechanism. On an industrial scale, lactide-to-PLA production is integrated into large facilities, such as NatureWorks' , plant with a capacity of 150,000 metric tons per year, supplemented by a new 75,000 metric tons per year site in slated to begin operations in 2025. This expansion reflects growing demand, with global PLA market volume projected at approximately 570,000 metric tons in 2025, primarily driven by packaging films and filaments that leverage PLA's renewability and printability. Production costs for PLA via this route range from $2.20 to $2.90 per kg, influenced by feedstock prices and regional energy inputs, positioning it competitively against petroleum-based alternatives in high-volume sectors.

Biomedical and Other Uses

Lactide-based copolymers, particularly formed by copolymerization of lactide with glycolide, are extensively utilized in biomedical fields such as , surgical sutures, and implantable devices. These materials enable sustained release of therapeutics and provide temporary structural support in the body, degrading into non-toxic byproducts like lactic and glycolic acids. has been approved by the U.S. Food and Drug Administration (FDA) since 1989 for therapeutic applications, owing to its and tunable degradation profile. In , scaffolds fabricated from lactide-derived poly() () support bone regeneration by mimicking the and promoting and . These scaffolds exhibit controlled rates spanning months to years, which can be adjusted via molecular weight and composition to align with the pace of new formation. For instance, scaffolds integrated with bioactive fillers enhance osteogenesis while maintaining mechanical integrity during the healing process. Lactide also finds application as a monomer in non-biomedical sectors, including the production of adhesives and protective coatings where lactide-based polyesters offer superior mechanical strength, rheological behavior, and corrosion resistance compared to conventional resins. Additionally, lactide enables practices through , where post-consumer is depolymerized back to high-purity lactide for repolymerization, reducing reliance on virgin feedstocks. Emerging uses of lactide-based materials include nanoparticles for targeted cancer therapy, which encapsulate chemotherapeutic agents to improve , minimize systemic , and enable site-specific delivery via enhanced permeability and retention effects in tumors. Furthermore, blends with other , such as poly(hydroxybutyrate), are incorporated into films to enhance gas barrier properties and extend without compromising biodegradability. The specialized nature of these biomedical and other applications is constrained by the high production costs associated with high-purity lactide required for medical-grade materials, limiting their scale to niche markets relative to bulk uses.

Safety and Environmental Impact

Toxicity and Handling Precautions

Lactide demonstrates low acute oral , with an LD50 exceeding 5,000 mg/kg in rats, indicating minimal risk from single ingestions. Dermal is similarly low, with an LD50 greater than 2,000 mg/kg in rabbits. The compound is not classified as a , and no reproductive or developmental has been identified in available assessments. Irritation potential varies by source: skin irritation is generally mild or none (per 404 rabbit studies), though hydrolysis to may cause irritation in some cases. For eyes, classifications range from moderate irritation ( 405, Draize scores indicating redness, swelling, and temporary discomfort; Category 2) to serious damage (Category 1, potential corneal injury). Inhalation of dust or mist may irritate the , causing coughing or , though specific LC50 data are unavailable and no (TLV) has been established by ACGIH. Handling in well-ventilated areas is recommended to minimize exposure. Safe handling requires , including chemical-resistant gloves, safety goggles, and protective clothing. Storage should occur in a cool, dry place at 2-8°C under to prevent to . No specific OSHA permissible exposure limits exist for lactide, but it is registered under REACH (EC number 202-468-3 for dilactide). Ingestion may cause adverse gastrointestinal and potential renal effects based on ; medical attention is advised. Chronic exposure data are limited, but hydrolysis to suggests low long-term risks, similar to .

Biodegradability and Sustainability

Lactide, the cyclic dimer of , readily undergoes in aqueous environments to yield monomers, which are naturally metabolized by microorganisms. In the context of () produced via of lactide, biodegradability primarily occurs through hydrolytic chain scission followed by enzymatic degradation. Under industrial composting conditions at thermophilic temperatures of 50–60 °C, hydrolyzes and biodegrades over 3–6 months, achieving near-complete mineralization to , , and via microbial action. In soil and environments, enzymatic breakdown by and fungi proceeds more slowly, often requiring elevated temperatures or specific microbial consortia for efficient degradation. The environmental impact of lactide-derived PLA is generally lower than that of petroleum-based plastics like (), owing to its renewable feedstocks such as . Life cycle assessments indicate that PLA production emits approximately 0.5–2.0 kg CO₂ equivalents per kg (cradle-to-gate, varying with biogenic carbon credits), compared to 2–3 kg CO₂ equivalents per kg for , representing a 30–50% reduction in . However, if not managed through industrial composting, PLA can persist in marine environments for extended periods, potentially fragmenting into that pose risks to ecosystems due to slow hydrolytic at ambient temperatures. Sustainability efforts for lactide-based materials emphasize closed-loop , where of via thermolysis regenerates lactide with recovery rates up to 90–98%, enabling high-purity reuse without loss of material quality. Comprehensive analyses confirm that such pathways further reduce emissions by 30–50% relative to virgin fossil-based plastics, promoting a . Despite these advantages, challenges include the energy-intensive nature of lactide purification and , which can offset some renewable benefits, as well as significant demands for corn feedstock , potentially competing with food production. Additionally, effective end-of-life management requires access to specialized composting facilities, limiting widespread in regions without . Looking ahead, emerging bio-based alternatives, such as lactic acid production from microalgal , aim to enhance circularity by reducing reliance on and improving overall metrics through lower water and inputs compared to corn-derived processes.

References

  1. [1]
  2. [2]
    Lactide: Production Routes, Properties, and Applications - PMC - NIH
    The lactide is a cyclic ester formed by two lactic acid molecules obtained via lactic acid polycondensation, followed by depolymerization of the PLA prepolymer ...
  3. [3]
    Lactic Acid Derivative - an overview | ScienceDirect Topics
    Structures of L- and D-lactic acids. Cyclization of two molecules of the corresponding lactic acid gives its cyclic dimer, called lactide. Thus, there are ...
  4. [4]
    Biodegradable Polylactic Acid and Its Composites - MDPI
    Jul 19, 2023 · Polylactic acid (PLA) is a biodegradable polyester polymer that is produced from renewable resources, such as corn or other carbohydrate ...
  5. [5]
    A review on biodegradable polylactic acid (PLA) production from ...
    Through the fermentation of renewable resources like sugarcane, starch and maize, LA is produced. PLA is a widely used material in human bodies since it is non- ...
  6. [6]
    Poly(lactic Acid): A Versatile Biobased Polymer for the Future with ...
    This review summarizes monomer synthesis of lactic acid and lactide, new synthesizing methods (melt polycondensation and ring opening polymerization) in PLA ...
  7. [7]
    PRODUCTION AND PURIFICATION OF LACTIC ACID AND LACTIDE
    Jun 3, 2022 · Pelouze , J. Chem. Pharm. 1845 , 53 , 112 – 124 . 10.1002/jlac ... Poly(Lactic Acid): Synthesis, Structures, Properties, Processing ...
  8. [8]
    Polymers Based on PLA from Synthesis Using D,L-Lactic Acid (or ...
    In 1954, a lactide purification method was developed by DuPont and made it possible to obtain high molar mass (HMW) PLA [70,72]. However, it was from 1966 ...
  9. [9]
    [PDF] polylactide (PLA) production - NatureWorks LLC
    In November 2001, NatureWorks started the production of PLA in its 140,000-tonne-per-year manufacturing facility in Blair, Nebraska. One year later ...Missing: 1990s | Show results with:1990s
  10. [10]
    Chemical Recycling of End‐of‐Life Poly(lactide) via Zinc‐Catalyzed ...
    Nov 2, 2020 · The chemical recycling of poly(lactide) was investigated based on depolymerization and polymerization processes.
  11. [11]
    The Chemical Recycling of PLA: A Review - MDPI
    This review covers the different methods of PLA chemical recycling, highlighting recent trends and advances in the area.
  12. [12]
    Lactide
    Insufficient relevant content. The provided text only includes a PubChem logo, a JavaScript requirement notice, and minimal structural information (Lactide | C6H8O4 | CID 7272). No IUPAC name, synonyms, molecular weight, physical description, melting point, or safety/use information is present.
  13. [13]
    Crystal and molecular structures of glycolide and lactide
    Aug 7, 2025 · The X-ray crystal structures of glycolide (1,4-dioxane-2,5-dione) and L(-)-and DL-lactide (3,6-dimethyl-1,4-dioxane-2,5-dione) are reported.
  14. [14]
    Direction of the dipole moment in the ester group - IBM Research
    May 1, 2002 · The moment of the ester group is directed (in the positive sense) at an angle τE = 123 ± 3° from the R-CO axis in both aromatic and aliphatic esters.Missing: lactide | Show results with:lactide<|separator|>
  15. [15]
    Direct Synthesis of Lactide from Lactic Acid by Sn-beta Zeolite
    In this study, we propose a continuous lactide synthesis process ... Mechanistic studies using isotopically labeled (2H, 13C) glucose reactants ...
  16. [16]
    L-Lactide(4511-42-6) 13C NMR spectrum - ChemicalBook
    L-Lactide(4511-42-6) 13C NMR. Product NameL-Lactide. CAS4511-42-6. Molecular FormulaC6H8O4. Molecular ...
  17. [17]
  18. [18]
    Determination and Correlation of Solubility Data and Dissolution ...
    Dec 6, 2012 · It was found that l-lactide is sparingly soluble in ethanol, isopropanol, methanol, and methylbenzene while it has high solubility in ethyl acetate and acetone.Introduction · Experimental Section · Results and Discussion · Conclusions
  19. [19]
    PLA architectures: the role of branching - RSC Publishing
    Dec 2, 2014 · Lactide contains two stereocentres, which lead to three optically active stereoisomers: L,L-lactide (or L-lactide); D,L-lactide (or meso-lactide) ...
  20. [20]
    Synthesis of L-Lactide from Lactic Acid and Production of PLA Pellets
    Feb 25, 2024 · The melting point of L-lactide of the highest purification degree was 96 °C, while those of the meso-lactide and racemic mixture of L- and D- ...<|control11|><|separator|>
  21. [21]
    Meso-lactide, processes for preparing it and polymers and ...
    The invention relates to a meso-lactide having a melting point of 52.8 DEG C., processes for its preparation, and its use in the preparation of resorbable ...
  22. [22]
    Determination of D-lactide content in lactide stereoisomeric mixture ...
    Mar 1, 2017 · ... lactide at 20 °C, the absolute error caused by the variation in temperature of 20±15 °C was not more than 0.2 and 0.7% in DCM and toluene ...
  23. [23]
    Poly(L-lactide) Epimerization and Chain Scission in the Presence of ...
    Jun 15, 2022 · Base-catalyzed hydrolysis mechanism proposed for the chain scission in the presence of phosphazene. Acid–base conjugates are known as catalysts ...
  24. [24]
    A greener process for poly-L-lactic acid production and chemical ...
    Feb 21, 2025 · Since epimerization was detected in several samples, their specific optical rotation was measured to determine the d-isomer percentage ...Missing: configurations | Show results with:configurations
  25. [25]
    None
    ### Method for Lactide Synthesis from L-Lactic Acid Using SnCl₂
  26. [26]
    Lactide Synthesis Using ZnO Aqueous Nanoparticles as Catalysts
    May 9, 2023 · The lactide synthesis yielded 77–80%, and the stability of the synthesized lactide was confirmed through a second purity determination after ...
  27. [27]
    Continuous process for manufacture of a purified lactide
    The material was heated to 200° C. and the pressure reduced to 10 mm Hg. The distillate was partially condensed to form a product fraction and the remaining ...
  28. [28]
    Ingeo Technology - NatureWorks LLC
    A proprietary two-step process transforms lactic acid molecules into rings of lactide, which is a valuable chemical on its own and the core of our customizable ...Missing: details | Show results with:details
  29. [29]
    US6326458B1 - Continuous process for the manufacture of lactide ...
    A process for the continuous production of substantially purified lactide and lactide polymers from lactic acid or an ester of lactic acid.
  30. [30]
    Microwave‐Assisted Synthesis of Poly(L‐lactic acid) via Direct Melt ...
    Dec 4, 2009 · A microwave-assisted method of synthesizing high-molecular-weight PLA using SSA as green catalyst was developed. Yellowish PLA with equation ...
  31. [31]
    Cost competitiveness of sustainable bioplastic feedstocks – A Monte ...
    The minimum selling prices for one t of LA, lactide and PLA are USD 943, USD 2,073 and USD 3,330, respectively. Sanaei and Stuart (2018) investigate the costs ...
  32. [32]
    L-lactide Market Size, Share, Trends, & Industry Forecast - 2031
    Feb 16, 2024 · The L-lactide market was valued at US$ 1.3 Bn in 2022; It is estimated to grow at a CAGR of 14.3% from 2023 to 2031 and reach US$ 4.3 Bn by ...Missing: 2020s | Show results with:2020s
  33. [33]
    Polylactic Acid Market Size, Growth Projections, Key Drivers, and ...
    May 3, 2023 · In 2021, the global Polylactic acid production capacity was estimated to be about 670 kilo tons, with global Polylactic acid demand of 428 kilo ...
  34. [34]
  35. [35]
    Ring opening polymerization of lactide: kinetics and modeling
    The main advantage of ROP is its living nature and the strict control of the polydispersity index (PDI). The livingness being a prominent feature of ROP ...
  36. [36]
    [PDF] Ring Opening Polymerization of Lactide for The synthesis of Poly ...
    Mar 2, 2006 · The ROP of lactide is thermodynamically driven by the relief of angle strain and switching from (E) to (Z) ester conformation upon ring opening.
  37. [37]
  38. [38]
    Biocompatible Catalysts for Lactide Polymerization—Catalyst ...
    Jul 13, 2020 · The activities of obtaining biocompatible catalysts of lactide polymerization were tested. The most active catalyst was zinc 2-ethylhexanoate (ZnOct 2 ).
  39. [39]
  40. [40]
    Enzymatic Ring-Opening Polymerization (ROP) of Polylactones - NIH
    ... opening polymerization (ROP) reaction; this type of enzymatic process is very sensitive to reaction conditions such as solvents, water content and temperature.
  41. [41]
    Are Well Performing Catalysts for the Ring Opening Polymerization ...
    Jun 8, 2022 · We report here the synthesis of two Zn(II) complexes supported by bulky monoanionic guanidinate ligands, which have been tested as catalysts for ...Introduction · Results and Discussion · Supporting Information · References
  42. [42]
    [PDF] Organocatalytic ring-opening polymerization of l- lactide in bulk - Pure
    Possible side reactions during the ROP of. LLA in bulk: (b) Base-catalyzed racemization of LLA via reversible deprotonation of the α-CH group with intermediate ...
  43. [43]
    How To Dry Lactide: The Complete Guide - Polylactide
    Lactide needs to be dry before polymerization to avoid hydrolysis and degradation. Some of the drying methods for lactide are: Vacuum drying: This method ...
  44. [44]
  45. [45]
    Depolymerization and Re/Upcycling of Biodegradable PLA Plastics
    Mar 13, 2024 · This review comprehensively outlines the current landscape of PLA recycling methods, emphasizing the advantages and significance of chemical re/upcycling.Figure 5 · 2.3. Alcoholysis · Figure 7
  46. [46]
    lactic acid)/Poly( D -lactic acid) Stereocomplexation and its Influence ...
    Sep 26, 2025 · Stereocomplexation between PLLA and PDLA is influenced by both temperature and time. •. The PLA stereocomplex promotes self-nucleation in PDLLA ...
  47. [47]
    About NatureWorks
    Our manufacturing facility, located in Blair, Nebraska, USA, has a name plate capacity of 330 million pounds (150,000 metric tons) of Ingeo biopolymer. Our ...Career Opportunities · Locations · News · Ingeo News
  48. [48]
    NatureWorks' Ingeo PLA Manufacturing Expansion Attracts Record ...
    May 15, 2024 · ... PLA manufacturing facility in Thailand which is on track to open in 2025. ... capacity of 75,000 tons of Ingeo biopolymer, it will produce ...
  49. [49]
    Polylactic Acid (PLA) Market Size & Share Growth 2034
    The global polylactic acid (PLA) market by volume was 476.79 KMT in 2024. It is estimated to grow at a CAGR of 20.60% from 2025 to 2034 to reach a volume of ...
  50. [50]
    Poly(lactic acid)/poly(lactic-co-glycolic acid)-based microparticles
    Jun 11, 2019 · Poly(glycolic acid), poly(lactic acid) and poly(lactic-co-glycolic acid) were approved by the United States Food and Drug Administration (FDA)
  51. [51]
    Poly(lactic-co-glycolic acid) microsphere production based on ... - NIH
    As such, PLGA has been approved by the Food and Drug Administration (FDA), as well as the European Medicines Agency (EMA), for orthopedics fixation, medical ...
  52. [52]
    [PDF] Medical Device Material Safety Summaries - ECRI Reports - FDA
    Nov 9, 2020 · There was no sign of exudate or ulceration in mice with PLGA plus medication. A high burst release percentage was observed for the implants with ...
  53. [53]
    PLA Scaffold Degradation & Acceleration for Bone Regeneration
    Nov 15, 2021 · However, PLA exhibits a very slow degradation rate, hindering the regeneration of new bone. In this study, the degradation mechanisms of PLA ...
  54. [54]
    PLLA Biomaterials for Regenerative Medicine: Processing & Apps
    Mar 14, 2022 · Their results showed that PLLA/PLA hybrid scaffolds had a faster degradation rate than pure PLLA and a lower mechanical strength at a higher ...
  55. [55]
    Technological performance of formaldehyde-free adhesive ...
    Main applications are in paints, coatings, adhesives, and sealants. The ... lactide in industrial production. Applications comprise bottles, cups ...
  56. [56]
    Lactide-based polyester as coating with improved properties
    Dec 10, 2021 · It revealed that lactide-based polyester resin had better mechanical, rheological, and anticorrosive properties than simple polyester resins.
  57. [57]
    PLGA-based nanoparticles for the treatment of cancer - AAPS Open
    Aug 1, 2022 · PLGA-based NPs can be used for a variety of cancer therapies including tumor-targeted drug delivery, gene therapy, hyperthermia, and photodynamic therapy.
  58. [58]
    Biopolymer Blends of Poly(lactic acid) and Poly(hydroxybutyrate ...
    Aug 16, 2022 · This study shows that the addition ChNCs in PLA:PHB can be a possible way to reach suitable gas barrier properties for food packaging films.
  59. [59]
    Poly(lactic acid) and Its Blends for Packaging Application: A Review
    Poly(lactic acid) (PLA) is a biodegradable aliphatic polyester produced from 100% renewable plant resources and plays a key role in the biopolymer market.
  60. [60]
    Applications of PLA in modern medicine - ScienceDirect.com
    This synthesis procedure requires high purity lactide and heavy metal catalysts which increase the cost of production. The company Mitsui Toatsu Chemicals ...Applications Of Pla In... · 2. Applications · 2.1. Tissue Engineering
  61. [61]
    A Comparative Review on Biodegradation of Poly(Lactic Acid) in ...
    From a biodegradation viewpoint, PLA demonstrates biodegradability in compost, wastewater, soil, under accelerated landfill conditions, and in water in ...
  62. [62]
    Accelerating the Biodegradation of Poly(lactic acid) through the ...
    A key characteristic of industrial composting conditions is temperature; thermophilic temperatures are critical to ensuring rapid degradation of PLA, which is ...
  63. [63]
    (PDF) Hydrolysis and Biodegradation of Poly(lactic acid)
    Aug 5, 2025 · Hydrolytic pretreatment, involving exposure to moisture or water, especially at elevated temperatures, can initiate cleavage of ester bonds in ...
  64. [64]
    Pivotal role of polylactide in carbon emission reduction: A ...
    May 5, 2024 · Specifically, PLA outperforms PET in terms of climate change, emitting 22% less GHGs, but falls short of PP's overall environmental impact.
  65. [65]
    Polylactic acid synthesis, biodegradability, conversion to ...
    Jan 25, 2023 · Here we review the polylactic acid plastic with focus on synthesis, biodegradability tuning, environmental conversion to microplastics, and impact on microbes.
  66. [66]
    Efficient chemical recycling of poly(L-lactic acid) via either ...
    Chemical recycling of poly(L-lactide) was achieved by properly designed Zn(II) catalysts. · Up to 99 % methyl lactate results by treating the polymer samples in ...
  67. [67]
    Characteristics of the steam degradation of poly(lactic acid ... - Nature
    Jan 24, 2024 · The total recovery rate was approximately 90%, and the remaining 10% comprised mainly highly volatile compounds, such as acetaldehyde, which ...
  68. [68]
    Environmental footprint of polylactic acid production utilizing cane ...
    Mar 10, 2025 · This study provides a comparative life cycle assessment (LCA) of PLA production from cane-sugar and microalgal biomass, analyzing contributions across ...