Open Access Article
Sweety Sharmaab,
Shantanu B. Sathayeac,
Yeddula Nikhileshwar Reddya,
Jayeeta Bhaumik
a and
Sudhir Pratap Singh
*ac
aCenter of Innovative and Applied Bioprocessing (BRIC-NABI), Sector 81, SAS Nagar, Mohali, India
bIndian Institute of Science Education and Research, Mohali, India
cDepartment of Industrial Biotechnology, Gujarat Biotechnology University, GIFT City, Gandhinagar, Gujarat 382355, India. E-mail: sudhir.singh@gbu.edu.in
First published on 5th January 2026
D-Allose is a rare monosaccharide and a low-calorie sweetener found in tiny amounts in nature with potential health benefits. Its scarcity in nature and high production cost pose a challenge to industry. The enzymatic route for D-allose production using L-rhamnose isomerase (L-RIM) is a promising and sustainable approach. The sensitivity of the enzyme to pH and temperature limits its application. In this study, we have developed manganese-based hybrid micro-structures (L-RIMnNFs) that offer robustness and recyclability to the enzymatic process of D-allose biosynthesis. The hybrid of organic and inorganic microstructure attains flower-like morphology, with each petal in the nanometer dimensions as visualized under electron microscopy. Molecular dynamics (MD) simulations were performed to understand the formation of these flower-like structures. These nanoflowers (NFs) exhibited better kinetic properties and reusability, retaining almost >60% activity even after the 6th cycle of reaction. So far, this study is the first attempt to immobilize L-rhamnose isomerase enzyme onto hybrid nanoflowers (hNFs) and to conduct simulation studies thereof. It provides insights into the improved robustness and durability of L-RIM after immobilization with a turnover number of 679.9 s−1. These NFs can be developed into biocatalytic platforms for the production of rare sugars.
L-Rhamnose isomerase is a potential enzyme for D-allose production.10 It is a multi-substrate enzyme with the highest specificity for L-rhamnose. Though this enzyme is capable of catalyzing the isomerization of many other monosaccharides like D-glucose, L-glucose, D-galactose, D-mannose, it has been well characterized for D-allose production.11 Generally, biocatalysts offer high specificity, low toxicity, high activity and high product yield under mild conditions. However, many enzymatic processes face the disadvantages associated with reusability and the functionality of the biocatalyst under harsh industrial conditions. Immobilization of enzyme molecules to a suitable matrix makes the bio-catalytic process robust and economical, conferring elevated stability and multiple catalytic cycles.12,13 Many reports have documented that immobilization leads to a better kinetics of the enzyme.14,15 Immobilization of L-rhamnose isomerase has been demonstrated on Chitopearl beads,16 anion exchange resins,17 Duolite beads,18 and as cross-linked clusters.19 In the aforementioned studies, enzyme immobilization has been devised on organic, inorganic, and mixed matrices, following the approaches of adsorption, entrapment or encapsulation and covalent modification.15,20 Immobilization events may alter the active site of the enzyme, leading to decrease in activity and increased mass transfer.21 The beads or resin carriers as immobilization platform are formed by conventional method i.e., adsorption, physical entrapment or covalent binding for the enzyme immobilization. The conventional immobilization suffers with enzyme leaching and/or diminished activity of the enzyme, in turn decreasing the loading capacity and catalytic efficiency. Further, covalent immobilization may alter the active site and often catalytic performance of the enzyme is compromised.22 Therefore, there is a need to follow alternative immobilization approaches for the improvement in bio-catalytic process without compromising the activity.23 To tackle this issue, preparation of self-assembling protein structures are being explored. The bottom-up approach to form ordered nano- or micro-scale structures improves enzyme activity by providing reduced mass transfer resistance and diffusional constraints. These assemblies provide accessible microenvironments while preserving the native conformational state of the protein as compared to conventional immobilization methods.12
Nano-flowers are formed by self-assembly of protein molecules in the presence of metal and phosphate ions, preserving the native state of enzyme, reducing the usage of expensive conjugating agents, and in these structural conjugates leaching of enzymes is largely avoided. Interestingly, nanostructure allow a greater exposure to the enzymatic active site for proper channelling and increased accessibility of the substrate.24,25
Different 3D architectures for enzyme immobilization have innumerable applications due to additional benefits of decreased size, increased surface area and multiple interfaces. The structures formed by self-assembly have drawn attention of researchers in recent years due to multitude of advantages, such as easier method of preparation, enhanced selectivity, activity and stability of the enzymes.12,26–29 The hybrid organic–inorganic flower-like structures, known as nanoflowers (NFs), are quite popular these days in the enzyme immobilization world.30 These structures were first reported by Ge et al. 2012, where they synthesized different kinds of NFs by using copper phosphate (Cu3(PO4)2) and a variety of proteins.31 These structures were formed by self-assembly between the amine/carboxyl group of protein (organic component) and metal phosphates (inorganic component).32 The synthesis takes place mainly in three steps: (1) nucleation: formation of primary crystals of metal and phosphate ions, (2) anisotropic growth: formation of coordination bonds between primary crystals and amino/carboxyl groups of side chains of amino acids of the enzyme, and (3) flower-like morphology: aggregation of petal-like nano-structures to form complete flower, where protein component acts as “adhesive” for sustaining such morphology. For metal dependent enzymes, it is an advantage to use inorganic component containing the desirable metal that acts as cofactor, enhancing the catalytic activity and structural integrity of the enzyme.
In the present investigation, we have immobilized L-rhamnose isomerase enzyme as an organic component into self-assembled flower-like structures, using manganese as inorganic component, to form a hybrid system. Different characterisation studies were performed to elucidate the structural and physical aspects of these structures. A simulation-based approach was used to get insights into possible mechanism involved in formation of these structures. Biochemical characterization of the NF structures was conducted to study catalytic efficiency, specificity and reusability of enzyme. This combinatorial platform is presumably a smart solution for the industrial production of rare sugars of high-value, such as D-allose.
The qualitative and quantitative analysis of loaded protein was performed by recording spectrophotometric absorbance and Bradford assay (Sigma-Aldrich, St. Louis, Missouri) by using bovine serum albumin (BSA) as a standard. Detailed method is provided in SI Section 2.2 and Fig. S2.
For functional group analyses, L-RIMnNFs samples were dried using lyophilyzer for 24 h. Then, each sample was applied onto the surface provided for sample analysis. The structural details were determined by Fourier transform infrared spectrophotometry (FT-IR) in the spectral range of 4000 to 400 cm−1 [attenuated total reflection (ATR) mode].
The crystallinity of the lyophilized hNFs along with other controls was determined by powder X-ray diffraction crystallography (XRD). The Cu Kα radiations (λ = 1.540593) were recorded at 40 kV voltage of diffractometer and in a scan range of 5–80 θ with a scan speed of 5° min−1. To validate the thermal stability of NFs, differential scanning calorimetry (DSC)/thermal gravimetric analysis (TGA) was done using ∼7 mg of freeze-dried sample. The analysis was carried out in the range of 30 °C to 600 °C, with a consistent heating rate of 15 °C min−1.
The Brunauer–Emmett–Teller (BET) analysis was carried out to measure the specific surface area (BET-SSA) and porosity of the synthesised nanocomposites. The synthesised NFs were degassed at 100 °C for around 2 h prior to analysis.
The size distribution of NFs was determined by DLS. The surface charge distribution of protein–metal complex was determined by measuring zeta potential of the NFs. The association of protein component with metal phosphates in the NFs and the uniformity of distribution were determined by fluorescent dye (FITC) tagging. For this, 50 µl of FITC (1 mg ml−1 in DMSO) was added to the enzyme followed by incubation at 4 °C with slow shaking for 8–12 hours. The tagged protein was then used for the NF synthesis and later monitored by confocal laser scanning microscopy (CLSM).
000 RPM for 20 min, followed by re-suspension in 50 mM HEPES buffer. The prepared NFs were stored at 4 °C till further use.
:
50
:
50, respectively. The monomeric protein structure of L-RIM was obtained from the AlphaFold2 structure prediction tool. The simulation was performed using the tool, GROMACS 2020.1.34 The simulations used the CHARMM36
35 force field. A dodecahedron-shaped simulation box was created with a 1 nm periodic boundary condition and filled with water molecules as the solvent. The total charge on the system was neutralized using Na+ or Cl− ions. The energy of the system was minimized using the steepest descent minimization algorithm.36 The systems were equilibrated under NVT (constant number of particles, volume, and temperature) for 2000 ps, followed by NPT (constant number of particles, pressure, and temperature). After the completion of both equilibration steps, MD run was performed for 100 ns. The LINCS algorithm was used for the covalent bond constraints. The electrostatic interactions were treated with the Particle Mesh Ewald (PME) method.37,38 The designed systems were simulated at two different temperature points, 4 °C and 25 °C, to evaluate the effect of temperature on the self-assembly of metal–protein structure. A protein simulation system, without any addition of the metal ion, was used as a control system (Table S1).
Each trajectory after the completion of MD run was analyzed using the Gromacs utilities. The root mean square deviation (RMSD), solvent-accessible surface area (SASA), and radius of gyration (Rg) was computed by using ‘gmx rms,’ ‘gmx sasa,’ and ‘gmx gyrate’ Gromacs utilities, respectively.39,40 The XMgrace tool was used to prepare the graphs.41
000 RPM/25 °C for 20 min to separate the NFs and stop the reaction. The separated supernatant was further boiled at 100 °C for 5 min to assure complete denaturation of the leached enzyme, if any. The mixture was further centrifuged and filtered through 0.2 µm filter. The filtered samples were diluted in 50% acetonitrile: water followed by analysis in HPLC (Agilent technologies) equipped with Zorbax-NH2 column and refractive index detector (RID). The column and the detector temperature were 40 °C and 45 °C, respectively. The mobile phase was acetonitrile and MilliQ water (mixed at 65
:
35 ratio), with a flow rate of 1 ml min−1.
SEM images of different stages of NF synthesis, and the effect of trypsin, EDTA, different enzyme concentrations, temperature and incubation period on NF's formation have been elaborated in SI Section 2.3. The optimization of NF synthesis was done by using different enzyme and/or metal salt concentrations and incubated for different time periods (Section S2.2, Fig. S3 and S4). The NF of desirable morphology and activity could be obtained by using 0.2 mg ml−1 enzyme, pH 7.4, and incubation at 4 °C for 48 h. The SEM and TEM analysis depicted the flower-like morphology of the nano-clusters (Fig. 2A–D).
The average RMSD value of the ‘Test_4’ system was found lower as compared to that of ‘Test_25’, signifying higher conformational stability of protein at 4 °C. The results are in agreement with the experimental data (Fig. S3 and S4) for the synthesis of the NFs.
The simulation systems, ‘Test_4’ and ‘Control_4’ (simulation of L-RIM only at 4 °C), were compared to investigate the possible impact of Mn2+ and PO42− ions on the synthesis of NFs at 4 °C. The RMSD trajectory analysis of ‘Test_4’ system was observed to attain a plateau at around 30 ns onwards, with minor fluctuations. This indicates that the protein complex system has achieved equilibrium for the synthesis of the hybrid nanoflowers. The snapshots at different time intervals (i.e., 0 ns, 30 ns, and 100 ns) displayed the clustering of Mn2+ and PO42− ions around the protein surface with time (Fig. 3A), which could be critical in the formation of self-assembled NF structures. This observation motivated us to explore the surface composition of the protein surface with respect to the solvent accessibility. The computed average SASA value for ‘Control_4’ is 233.07 ± 1.823 nm2, while for the ‘Test_4’ system, it was 230.84 ± 1.83 nm2, indicating relatively reduced solvent accessibility.42 Though there was a minute difference in the average SASA values, ‘Test_4’ system showed a pattern different from that of ‘Control_4’ in the simulation trajectory. ‘Test_4’ system showed a gradual decrease in the accessible area marking from 30
000 ps (i.e., 30 ns) onwards, in contrast to ‘Control_4’ system (Fig. 3B). This implies the gradual reduction of protein backbone accessibility to the solvent molecules in the case of ‘Test_4’ system. This is possibly due to the formation of cluster of ions surrounding the protein surface. The self-assembling of metal ions and protein leads to increase the compactness of the structure.43 The average radius of the gyration (Rg) of the ‘Control_4’ and ‘Test_4’ was noted to be 2.26 ± 0.01 nm and 2.24 ± 0.01 nm, respectively (Fig. 3C). The lower Rg in ‘Test_4’ system indicates increments in the compactness of the protein structure with time. Interestingly, protein-ion complexes showed a common pattern of surface metal ion clusters that can be visualized into 3 regions highlighted in Fig. 3D. In cluster I, the residues involved in binding with the phosphate and manganese ions were identified as GLU 20, ARG 21, GLU 386, LEU 383, ARG 367, ASP28, and ALA 24. Cluster II showed the metal binding residues as ARG 300 and ARG 297, while HIS 259, ASP 216, and GLU 213 were metal associated residues of cluster III. Besides these clustered residues, 4 residues were noticed to have metal interaction viz., ARG 95, ASP 279, ASP 290, and ARG 76 (Fig. 3D). The results endorsed the essentiality of Mn2+ and PO43− ions in stabilizing the protein structure and facilitating self-assembly.
As immobilization of enzyme can induce changes in the conformation of a protein,44 it is advisable to assess the possible changes in the protein sites critical for substrate binding and catalysis in nanoflower particles. The amino acid residues crucial for substrate binding (Glu233, Asp334, Asp267, and His294) in the protein-ion complex system (Fig. S5C) were found same as identified for the free form of the protein, L-RIM.33 In L-rhamnose isomerases, the evolutionarily conserved α1–β1 loop plays an important role in executing catalytic activity.45 The superimposition of α1–β1 loop in simulated L-RIMnNFs and L-RIM in its free form displayed the RMSD value of 0.8 nm. The results infer that the simulated L-RIMnNFs did not exhibit any significant change in the conformation of evolutionarily conserved α1–β1 loop (Fig. S5D). This signifies that the process involved in the synthesis of NFs should not exert any undesirable impact on the catalytic activity of L-RIM (Table S1 and Fig. S5B).
The presence of Mn and other elements indicated the prominent incorporation of protein component into NFs. The transmission electron microscopic (TEM) analysis presented the contrasting image of the petals, and the grooves formed by their assembly affirming the porosity of the structure (Fig. 2C and D). This gives the reason for high surface to volume ratio of these structures. TEM also confirmed the size range of 7–9 µm of NFs, while the petals were 400–500 nm in length. Similar images were reported by Rai et al. and Huang et al.50,51
For gaining the structural details of immobilized L-RIM and chemical composition of NFs, FTIR in ATR mode was performed in wavelength range of 4000 to 400 cm−1. The enzyme's FTIR spectra exhibited prominent absorption bands near 3267 cm−1, attributed to O–H and N–H stretching vibrations, and characteristic of hydroxyl and amine groups in protein. This band, albeit with reduced intensity, can be noted in the enzyme nanoflower, suggesting the presence and attachment of the enzyme in the nanoflower. The prominent peak between 1600 and 1700 cm−1 corresponds to the amide I band (primarily C
O stretching), while a peak around 1540 cm−1 is indicative of the amide II band (mainly N–H bending and C–N stretching).52 Furthermore, a distinct C–N (1250 cm−1) stretching vibration, characteristic of protein or enzyme amide linkages, was importantly observed in the pure enzyme spectrum. This band appeared with reduced intensity in both the bare nanoflowers and the enzyme-nanoflower composite, indicating successful enzyme immobilization on the Mn nanoflower surface through weak covalent or electrostatic interactions. A small band nearly at 1633 cm−1 indicates bending vibration of O–H groups in the adsorbed water molecules. The presence of these amide bands in the conjugated sample confirms that the enzyme retains its structural features upon conjugation, indicating minimal denaturation. A broad peak observed between 500 and 650 cm−1 corresponds to the Mn–O stretching vibrations, confirming the presence of manganese oxide in the nanoflower structure. Moreover, Mn nanoflowers exhibit eminent peaks around ∼1070 cm−1 and ∼950–900 cm−1 due to P–O asymmetric stretching (PO43−) and metal oxygen phosphate (Mn–O–P) vibrations. Additionally, the fingerprint region (500–1500 cm−1), which contains complex and characteristic vibrational modes, shows several retained features in the conjugated sample. This further supports that the enzyme maintains its structural integrity and potentially remains functionally active when immobilized on the nanoflower surface.
The absorption bands at 1152 cm−1 corresponds to stretching vibration of P
O bond, 1049 cm−1 to asymmetrical stretching of P–O bond, 989 cm−1 to stretching vibration of P–O bond, 624 cm−1 to bending vibration of P–O bond and 553 cm−1 to in-plane bending vibration of phosphate ion, indicating the presence of phosphate groups.53 Also, there was no significant peak shift in NF's spectra as compared to free enzyme control (Fig. 5A), indicating the self-assembled nature of synthesized nanostructures, possibly formed by coordination bond rather than covalent bonds.54
![]() | ||
| Fig. 5 Physical characterization of L-RIMnNFs. (A) FTIR analysis (B) XRD analysis, (C and D) TGA and DTG graphs (blue line: metal phosphate control; orange line: L-RIMnNFs.). | ||
The crystallinity of the nanoflowers was analyzed by powder X-ray diffraction (PXRD). The relative peaks obtained in XRD of NFs showed similar pattern of peaks (Fig. 5B), as previously reported for Mn3(PO4)2·3H2O in the Inorganic Crystal Structure Database (ICSD) (JCPDS no. 00-003-0426). Since proteins are amorphous, the additional peaks could be attributed to the presence of phosphate along with manganese crystals.45,49,51,55
The absorption and desorption studies were carried out by using the Brunauer–Emmett–Teller (BET) analysis for specific surface area (SSA) and pore size determination.46,56 Fig. S7A shows the N2 adsorption/desorption isotherm of enzyme-immobilized particles. The shape of the curve suggests a Type IV isotherm, characteristic of mesoporous materials with average pore size of 5.1 Å in Mn and Mn-enzyme complex. The sharp uptake near p/p0 ≈ 1.0 is indicative of capillary condensation in mesopores, supported by mesoporosity. This indicates a pore shape i.e., cylindrical or slit-like pores. The material exhibits mesoporosity with significant surface area and pore volume. On the other hand, the surface area of these self-assembled structures was found to be 58.89 m2 g−1. These findings confirm the formation of an assembly between the organic and inorganic components.57
The temperature and heat flow associated with the nanomaterial transitions as a function of time and temperature was done by differential scanning calorimetric (DSC) and thermal-gravimetric analysis (TGA). The weight loss in the case of MnPO4 crystals was observed in two stages. The first weight loss was noted at around 80 °C that could be due to loss of water molecules. Up to 300 °C the weight loss could be due to the denaturation of the enzyme. The major weight loss (40%) was observed at around 300 °C and higher temperatures, which could be attributed to decomposition of metal phosphate crystals. In case of enzyme immobilized crystals, the proportion of weight loss was greater at 300 °C and higher temperatures, which could be due to decomposition of the organic component (i.e., enzyme) of the nanoflower. The results validated the encapsulation of enzyme in manganese phosphate nano-crystals (Fig. 5C and D).49,58
The hydrodynamic size distribution of L-RIMnNFs suspended in buffer solution was determined by differential light scattering (DLS) (Fig. 6A). A uniform distribution of NFs was noted in the range of 700 to 900 nm. This size was close to the size determined by TEM analysis.59 The stability of the suspended NFs was determined by measuring the surface charge distribution i.e., zeta-potential of NFs. A shift in the zeta potential from −8.78 mV to +6.51 mV indicated loading of the enzyme onto the metal phosphate crystals forming aggregates in the form of NFs.60
![]() | ||
| Fig. 6 Physical characterization of L-RIMnNFs. (A) DLS, (B) UV-vis spectrophotometry, and (C) CLSM activity assay and product analysis of L-RIMn-NFs. | ||
The distribution of protein component in the L-RIMnNFs was analyzed by confocal laser scanning microscopy (CLSM). Fluorescently labelled protein emitted green spectra at around 525 nm. The NFs synthesized with labelled protein exhibited uniform distribution all over the nanoflower structure (Fig. 6C).54,61 The spectrophotometric analysis showed the highest peaks at 230 and 280 nm, with estimated protein concentration of 0.156 mg ml−1 and a calculated enzyme load of 78% (Fig. 6B).
The activity assay was performed under the optimized reaction conditions for determining the conversion efficiency of L-RIMnNFs (Fig. S9A). The immobilized enzyme showed improved conversion over free enzyme, with almost 1.5 folds higher activity than the free counterpart. Since metal ions boost the activity of the enzymes, a system where framework is composed of favourable metal ions further improves the activity of the enzyme in the NFs. Such enhancement in the activity of the enzyme upon immobilization onto NFs has also been reported earlier.50
Immobilized biocatalyst might face alteration in its pH activity profile.64 However, in the present study, the immobilization of enzyme did not lead to any major change in the optimum pH. However, the pH stability of the enzyme was greatly improved as compared to its free counterpart, with highest stability at pH 6.0 and 7.0. The reason behind this could be a better shielding of the active site of the enzyme in acidic or alkaline environment. The NFs exhibited substantial activity in the pH range of 5.0 to 10.0, achieving the maximum relative activity at neutral pH (Fig. 6B).65
| Kinetic parameter | Free enzyme | Immobilized enzyme |
|---|---|---|
| Km | 110 mM | 64.44 mM |
| kcat | 328.8 s−1 | 679.9 s−1 |
| kcat/Km | 2.981 M−1 s−1 | 10.55 M−1 s−1 |
| This journal is © The Royal Society of Chemistry 2026 |