Next Article in Journal
FYCO1 Peptide Analogs: Design and Characterization of Autophagy Inhibitors as Co-Adjuvants in Taxane Chemotherapy of Prostate Cancer
Previous Article in Journal
Colonizing Bacteria Aggravate Inflammation, Cytotoxicity and Immune Defense During Influenza A Virus Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Binding Capacity and Adsorption Stability of Uremic Metabolites to Albumin-Modified Magnetic Nanoparticles

1
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB T6G 2R3, Canada
2
The Metabolomics Innovation Centre, University of Alberta, Edmonton, AB T6G 2E9, Canada
3
Department of Biological Sciences, University of Alberta, Edmonton, AB T6G 2R3, Canada
4
Department of Medicine, Cumming School of Medicine, University of Calgary, Calgary, AB T2N 2T8, Canada
5
Department of Biomedical Engineering, University of Alberta, Edmonton, AB T6G 2R3, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(11), 5366; https://doi.org/10.3390/ijms26115366
Submission received: 23 April 2025 / Revised: 22 May 2025 / Accepted: 27 May 2025 / Published: 3 June 2025
(This article belongs to the Section Molecular Endocrinology and Metabolism)

Abstract

Kidney disease causes the retention of uremic metabolites in blood, which is associated with many comorbidities. Hemodialysis does not properly clear many metabolites, including large, middle-sized, and small protein-bound uremic toxins (PBUTs). Adsorption strategies for metabolite removal require the development of engineered adsorbents with tailored surfaces to increase the binding of desired metabolites. Albumin is uniquely positioned for modifying blood-contacting surfaces to absorb uremic metabolites, as it (i) minimizes non-specific protein adsorption and (ii) binds a range of molecules at Sudlow Sites I and II with different affinities. It is unknown if albumin-modified surfaces retain the adsorption qualities of solution-free albumin, namely, adsorption stability or specificity. Herein, albumin was covalently attached to iron oxide nanoparticles and characterized using multiple methods. Metabolite adsorption was conducted by incubating particles in a model solution of thirty-three uremic metabolites associated with kidney failure. Adsorption efficiency, selectivity, and stability were affected by albumin concentration and incubation time. Metabolite adsorption was found to change with time, and it was more effective on albumin-modified particles than unmodified controls. The findings outlined in this paper are crucial for the design of next-generation advanced blood-contacting materials to enhance dialysis and blood purification for patients with kidney disease.

1. Introduction

Chronic kidney disease (CKD) affects 8 to 16% of the global population and ranks among the top 10 causes of death worldwide [1,2]. As kidney dysfunction progresses, the unavoidable accumulation of uremic metabolites (UMs) occurs [3,4]. Although life-saving, current hemodialysis (HD) methods fail to effectively remove all UMs, particularly protein-bound uremic toxins (PBUTs) and those with large distribution volumes, which remain at high concentrations due to reabsorption or recycling [4,5]. This has been linked to diminished life expectancy, reduced quality of life, significant comorbidity burdens, and high healthcare costs [6,7,8]. Given the constraints of hemodialysis and the minimal advancements in the lives of patients, it is estimated that kidney disease will rank as the fifth largest cause of death worldwide by 2040, so there is an urgent necessity for new effective treatment modalities [9,10].
Emerging adsorbent-based strategies have the potential to be used in wearable devices with many ideal attributes, like continual clearance, less water waste, less plastic waste, and better patient quality of life and health outcomes [11,12]. Magnetic nanoparticles (MNPs) have many notable properties for adsorbing metabolites from the blood, including a high surface area that is easily modified, proven biocompatibility, benign biodegradability, can be captured using exogenous magnetic fields, and low cost [13,14,15,16]. Blood-contacting adsorbents must be hemocompatible whilst having a sustained and specific binding of metabolites [17]. Albumin has been shown to bind a variety of molecules, including UMs, through Sudlow Sites I and II [18,19,20]. It is thought that modifying MNPs with albumin will enhance MNP hemocompatibility, dispersion stability, and oxidation resistance while providing a matrix to adsorb UMs [16,21,22,23,24].
In general, a significant gap exists in the literature on the adsorption of UMs from complex solutions of metabolites, where most work primarily focuses on individual toxins in single buffer solutions, and limited studies have addressed the uptake of these toxins by albumin-modified surfaces [25]. However, a plethora of metabolites accumulate in kidney failure patients that are linked with poor patient outcomes: xanthine, kynurenine, putrescine, creatinine, quinolinic acid, and polyamines [26,27,28,29]. Our lab has modified MNPs with (methacryloyloxy)ethyl phosphorylcholine-co-β-cyclodextrin (p(MPC-co-PMβCD)) films that are particularly effective in adsorbing UMs, displaying notable selectivity and time-dependent interactions [30]. Although MPC is considered a hemocompatible coating, it is possible that albumin modification will provide an increased density of metabolite binding sites and enhanced hemocompatibility through a simple surface modification approach.
UM interactions with albumin films are underexplored. Building upon our previous work on albumin-modified magnetic nanoparticles that studied their hemocompatibility [24], albumin-modified MNPs formed using similar techniques were exposed to a model solution of 33 UMs (Table 1) to understand how surface properties and incubation time affect metabolite binding capacity, adsorption changes with time, and adsorption specificity. Albumin coupling to an amino-silane-modified MNP was conducted at two different albumin solution concentrations (0.2 and 2 mg/mL), and formed films were characterized using thermogravimetric analysis (TGA), X-ray photoelectron spectroscopy (XPS), and zeta potential. Changes in the secondary structure of albumin within the coating were characterized using Fourier-transform infrared spectroscopy (FTIR), particle size, and morphology, and film thickness was quantified using transmission electron microscopy (TEM). Quantitative mass spectroscopic analysis was used to determine the adsorbed uremic toxin profile. Surface properties and incubation time (from 1 h to 4 h) were major determinants of metabolite adsorption—adsorption was not constant with time and not purely driven by solution concentration. These findings are pivotal for developing advanced, next-generation materials that enhance dialysis and blood purification, offering personalized treatment options for kidney failure patients.

2. Results

2.1. Particle and Film Dimensions

MNPs were analyzed using TEM to determine their size, shape, and homogeneity, both before and after albumin modification (Figure 1). Bare MNPs were mostly uniform and spherical, with diameters ranging from 8 to 14 nm, averaging 10.6 ± 3.1 nm (n = 104; Figure 2a). Modifications with APTES (NH2(CH2)3Si(OC2H5)3) and glutaraldehyde showed minimal size or shape changes, consistent with previous studies (Figure 1b,c) [40]. The TEM images showed the formation of an albumin layer (appearing as a lighter outline around the particles) in Figure 1d,e. The average diameters of particles coated with albumin were 20.9 ± 5.2 nm at 0.2 mg/mL and 34.7 ± 3.2 nm at 2.0 mg/mL albumin concentrations (Figure 2b,c). The 0.2 mg/mL result was consistent with a thin albumin layer, as expected from the hydrodynamic radius of bovine serum albumin (BSA) [41]. At 2 mg/mL, particle sizes ranged from 27 to 42 nm, varying with the nanoparticle type and albumin attachment method, as previously reported [21,40,41]. Higher concentrations of albumin led to MNPs with a globular, uniformly spherical shape and the formation of larger aggregates. There was also some excess protein, indicated by a light, cloudy residue on the TEM images, which obscured the precise measurement of the albumin coating thickness, though a substantial, densely coupled protein layer was present. Origin 2024’s t-test was used to conduct statistical analysis. Calculated p-values for bare and MNPs-APTES-GA-BSA (0.2), bare and MNPs-APTES-GA-BSA (2) and MNPs-APTES-GA-BSA (0.2), and MNPs-APTES-GA-BSA (2) were p < 0.0001, p < 0.0001, and p < 0.001, respectively.
Selected area electron diffraction (SAED) analysis was performed on bare MNPs, MNPs-APTES-GA-BSA (0.2), and MNPs-APTES-GA-BSA (2) (Figure 1f–h). The results were consistent with the structure of magnetite, with the SAED images indicating the presence of the {311}, {400}, {422}, {511}, and {440} families of planes. The diffraction analysis of Alb-MNPs showed similar patterns to bare particles—more dots were visible due to the protein layer. SAED analysis could not be properly conducted for thicker albumin layers, as the amorphous protein interfered with the diffraction pattern.

2.2. Relative Mass of Alb-MNP Films

For each modification step, TGA was conducted and summarized for all types of nanoparticles (Figure 3). The mass reduction observed below 100 °C was due to the evaporation of bulk water, while that between 100 and 200 °C was due to vicinal water content, which was likely trapped within the particle matrix. Bare nanoparticles exhibited approximately 9.5% weight loss, signifying minimal water content and absence of deterioration [6]. Validation of every modification was confirmed by additional mass loss in every step. Upon addition of glutaraldehyde, the reduced mass loss at lower temperatures and increased mass loss at higher temperatures indicated enhanced thermal stability, which is in agreement with our previous studies on similar systems—illustrating consistent surface modification [24].
The deterioration of BSA in films was not influenced by the BSA concentration. Further, up to 100 °C mass loss was attributed to denaturation of BSA, and mass loss before 200 °C was attributed to water loss [24,42]. Furthermore, albumin decomposed in the temperature range of 250–400 °C. Up to 110 °C, a broad peak was observed, where an endothermic effect has been attributed to BSA denaturation [5,24,42]. A noticeable change occurred between 180 and 500 °C, which is related to the decomposition of BSA, glutaraldehyde, and amine functional group complex. We propose that the noticeable shift in the maximum decomposition rate with varying BSA concentration confirms the adsorption of BSA on the magnetic nanoparticles [42]. Upon increasing the BSA concentration (0.2 to 2 mg/mL), the peak intensity increased, which reflected the shift toward higher mass loss. Nanoparticles modified with a high BSA concentration (2 mg/mL) showed the highest mass loss between 200 and 380 °C, which was linked to C-O and C-C bond breakdown, confirming the high amount of BSA coating, which can be confirmed by TEM images.

2.3. Surface Charge Analysis

Zeta potential measurement revealed the significant variation on surface modification of MNPs. Bare, APTES, APTES-GA, APTES-GA-BSA (0.2), and APTES-GA-BSA (2) MNP systems exhibited zeta potentials of −27.6 ± 2, −18.8 ± 1, −8.5 ± 2, −20.16 ± 2, and −29.25 ± 2 mV, respectively (Figure 4), at pH 7, consistent with previously reported values on similar surfaces [21,24,43,44]. The successful modifications with APTES and glutaraldehyde were further confirmed thorough subsequent changes in surface charge. It was observed that the addition of albumin significantly reduced the zeta potential due to its negative surface charge at this solution pH [43]. An increased concentration of albumin solution led to a greater negative surface charge, which has been correlated with improved suspension stability [45].

2.4. Film Bonding Chemistry

An XPS survey scan of MNPs before and after modification was conducted (Figure 5). Previous studies indicated the absence of a satellite peak between Fe 2p1/2 and Fe 2p3/2 for Fe3O4, and our current investigation supported this finding [24,46,47]. The peak positions for Fe 2p3/2 and Fe 2p1/2 were at 710.6 and 724.1 eV, respectively. The peaks observed for Fe 3p3/2 at approximately 55 eV, O(1s) at 529.5 eV, and C (1s) at 283.5eV corresponded to the Fe-O bond characteristic of the Fe3O4 nanoparticles (as expected) [47]. The presence of nitrogen and silica peaks in the survey scan confirmed silanization. The N 1s peak in APTES and its modified surfaces formed from the convergence of two separate peaks of NH2 at 400.7 eV and NH3+ at 401.8 eV, and the Si 2p peak at 98.7 eV. Similar to previous literature, the elevated protonation level in the amine groups at the end of the APTES chain was confirmed by the high-resolution scan of the nitrogen peak (Figure 5, inset) [48]. A noticeable shift in the N 1s peak following the addition of glutaraldehyde indicated significant chemical modification and the formation of new bonds: C-N = C found at 399.05 eV assigned to the aldehyde group in glutaraldehyde and the amine group in APTES [49].
The successful film coating was further confirmed by the intensity shift in the O 1s, C 1s, and Fe 2p peaks. Initially, the major contribution to the peak, observed at 529.5 eV, originated from oxygen in the crystal structure, typical of oxide crystal formations. A hydroxyl functional group on the nanoparticle surface was also identified at 530.9 eV. Subsequent APTES modification shifted the hydroxyl peak to 531.6 eV due to covalent bonding with silicon on the surface. The N 1s peak increased with APTES, and the N 1s and C 1s peaks increased with albumin, suggesting the covering of the Fe surface and increased fraction of these components within the escape distance of the photoelectron. Moreover, albumin binding to the magnetic nanoparticle led to significant peak shifts attributed to various functional groups within the protein structure, including OH, O-C, O-H, O-N, and O-C-N.
XPS analysis was used to determine the chemical composition of samples (Table 2). The detection of nitrogen and silicon values validated the effective surface modification achieved by APTES. The values for nitrogen and silicon were very similar before adding albumin, which was expected, as APTES contains a 1:1 ratio of nitrogen and silicon, and glutaraldehyde includes neither. The atom % of iron decreased with each chemical addition, which was expected, as the coating thickness increased. With the addition of albumin from 0.2 to 2 mg/mL, the relative carbon content in the samples significantly increased for higher concentrations of BSA film, indicating a densely packed albumin layer. Significant carbonaceous contamination was present in the bare MNP sample. These values were similar to those previously reported for surface modification by APTES and GA on MNPs [48,49].

3. Discussion

3.1. Effect of Film Formation on Albumin Secondary Structure

Fourier-transform infrared spectroscopy (FTIR) was utilized to observe changes in the secondary structure of albumin to understand the extent of changes upon incorporation within the film. This was done to understand if denaturation was so extensive that the Sudlow Sites could not be maintained, thus altering UMs’ retention. It was confirmed that albumin was present on the MNP surface (Figure 6), as indicated by the characteristic amide-I (-NH2-) stretching peaks at 1649, 1648, and 1653 cm−1, as well as the amide-II (-NH-) bending peaks at 1533, 1539, and 1534 cm−1. The drop in intensity suggested a structural transition from α-helix to β-sheets and dehydrated β-turns, confirmed by the 1394, 1385, and 1396 cm−1 peaks. Alterations in dehydrated β-turns were identified using the bands at 1243, 1248, and 1246 cm−1. The presence of the Fe-O bond was also confirmed using the peaks at 556 and 634 cm−1. The percentage of secondary structural elements was quantified by fitting the amide-I band region with Gaussian curves and finding the multi-peak position using Fourier self-deconvolution (FSD) of the protein amide-I band. Multicomponent peak areas were quantified using Gaussian fitting in Origin 2024 (Figure 7).
Characterization of the secondary structures of BSA was conducted (Table 3) using the amide-I region, as determined using: (i) dehydrated β-turns, 1694 cm−1; (ii) β-sheet/turns, 1685–1663 cm−1; (iii) α-helix, 1660–1650 cm−1; (iv) random coil, 1648–1638 cm−1; (v) β-sheet, 1639–1621 cm−1; (vi) side chain, 1616–1600 cm−1 [50,51]. A variability of ~10% in secondary structures has been previously reported, with an α-helix content typically comprising 58–65% of the total secondary structure [52,53,54]. The α-helix content is essential, as Sudlow Sites I and II reside in helix-rich protein regions. The α-helix content for BSA has been shown to decrease upon film formation, and random coil content increased compared to when in solution [53,54]. Loss of helical structure was greater for MNPs-APTES-GA-BSA (0.2) than MNPs-APTES-GA-BSA (2), possibly due to the covalent bonding of albumin at the surface. Previous studies have indicated that the loss of α-helix structure was compensated by the emergence of beta structures due to the relative stability of beta structures compared to α-helices [50,53]. For MNPs-APTES-GA-BSA (2), the loss of β-sheets was more than that for MNPs-APTES-GA-BSA (0.2), while the major gain for MNPs-APTES-GA-BSA (2) was in β-turn, and for MNPs-APTES-GA-BSA (0.2) was in the random chain and dehydrated β-turns. This indicated that the albumin film was destabilized due to interaction with the MNP surface. However, the β-sheets remained low, indicating that no BSA aggregation occurred for MNPs-APTES-GA-BSA (0.2) [51,55]. Additionally, for MNPs-APTES-GA-BSA (2), the increased content of β-turns may prevent denaturation by inhibiting the interaction that exposes protein fragments susceptible to major conformational changes that lead to the unfolding of the protein [56]. It seems that changes in α-helix content transitioned to increased random coils and dehydrated β-turns for the MNP-APTES-GA-BSA (0.2) system, suggesting a hard protein corona formation with slight random coil formation. The higher-concentration system showed no dehydrated β-turn formation, suggesting a soft corona formation and only a moderate increase in β-sheet and β-turn composition. It seems that tethered albumin underwent a change in protein structure, but the protein remained intact and similar to its native form [52,57].

3.2. Uremic Metabolite Adsorbed Amounts

It was observed that all MNP systems adsorbed metabolites (Figure 8), but the quantity and types of metabolites were dependent upon the MNP surface and incubation time. The total adsorbed UMs for Alb-MNPs (0.2 mg/mL) were ~45 and ~44 nmol at 1 and 4 h of incubation, respectively. Alb-MNPs formed from 2 mg/mL films showed the highest adsorption at 1 h, followed by a decrease after 4 h from ~57 to ~41 nmol. For 1 and 4 h of incubation on bare MNPs and MNP-APTES-GA particles, the total adsorbed mols were ~37 to ~51 nmol and ~40 to ~49 nmol, respectively. In the case of bare particles (1 and 4 h) and MNPs-APTES-GA (1 and 4 h), an increase in incubation time yielded an average increase in total mols of adsorbed metabolites. On average, the total mols adsorbed decreased with the increasing incubation time for MNPs-APTES-GA-BSA (2), suggesting metabolites may be displaced with time in this system. That said, in all ANOVA tests using Tukey’s multiple comparison tests, the p-value for bare MNPs (1 h) vs. bare MNPs (4 h) was 0.046, bare MNPs (1 h) vs. MNPs-APTES-GA-BSA (2) for 1 h was 0.013, MNPs-APTES-GA (1 h) vs. MNPs-APTES-GA-BSA (2) for 1 h was 0.045, and MNPs-APTES-GA-BSA (2) for 1 h vs. MNPs-APTES-GA-BSA (2) for 4 h was 0.034, and others were not significantly different.

3.3. Uremic Metabolite Adsorption Selectivity

Adsorption selectivity was evaluated by normalizing the adsorbed UM (mol%) to that of the corresponding UM amount (mol%) in the original solution (Table 4). All MNPs had selectivity of 20 to 40 and 2 to 4 times greater than the original solution concentration for the neutral pyruvic acid and methylhistidine [58,59]. This was despite the low initial mol% of methylhistidine in the original solution. In contrast, none of the MNPs adsorbed detectable quantities of phenylacetic acid, even though it was 42 mol% of the initial solution. Similarly, L-asparagine, L-tyrosine, p-hydroxylphenylacetic acid, putrescine, homocysteine, xanthine, spermidine, spermine, argininic acid, and asymmetric dimethylarginine were not detected. Hippuric acid and creatinine, present in original solutions at 16 and 15 mol%, respectively, showed a significantly lower adsorbed mol%. An increased selectivity for 4-ethylphenyl sulfate and indoxyl sulfate was observed for most Alb-MNPs. Although, indoxyl sulfate and p-cresol sulfate binding with albumin was expected to be enhanced, while this was not observed for p-cresol sulfate under these conditions [19,20]. Both p-cresol sulfate and hippuric acid bind to Sudlow Site II, and their reduction may arise due to the denaturation of Sudlow Site II upon film formation or other compounds blocking access to this site [18,20]. Other than bare MNPs, the adsorption of indoxyl sulfate was up to 1.7 times that of the initial solution, increasing with the increased incubation time. These findings strongly demonstrated that individual metabolite adsorption was not solely dependent on their concentration in the incubation solution, as previously observed [30].
Uremic metabolite adsorption after 1 and 4 h of incubation was analyzed (Table 5) and showed that the interaction between metabolites and surfaces changed with time, a result considered holistically earlier in Figure 8 but now discussed on a per UM basis. Pyruvic acid was found to be highly adsorbed, which greatly affected the total adsorbed amounts of metabolites for each MNP system and the selectivity of the particles with time. Bare MNPs and MNP-APTES-GA particles exhibited a significant decrease in mol% for every compound except pyruvic acid and kynurenine, which showed an increase. Interestingly, MNP-APTES-GA-modified MNPs exhibited a decrease in mol% for most of the metabolites except for kynurenine, pyruvic acid, indoxyl sulfate, and 4-ethylphenyl sulfate, which had a substantial increase in mol% of 100, 18.6, 17.9, and 1.9, respectively, suggesting that this surface chemistry has a high selectivity for kynurenine, pyruvic acid, and indoxyl sulfate.
Conversely, Alb-MNPs had a significant increase in almost every compound except for pyruvic acid, 4-ethylphenyl sulfate, and kynurenine from 1 to 4 h of incubation. The latter showed drastic decreases of up to 420% between 1 and 4 h. This trend was magnified for all UMs adsorbed to MNP-APTES-GA-BSA (2 mg/mL) systems, suggesting that more albumin led to more binding than 0.2 mg/mL systems. These Alb-MNP results were directly opposite to the bare MNP and MNP-APTES-GA systems. Most notably, bare MNPs and MNP-APTES-GA showed a substantial increase in mol% of pyruvic acid from 1 to 4 h of incubation. In contrast, adding albumin yielded a significant decrease in the same timeframe, perhaps indicating that with increased time, pyruvic acid was being displaced from the surface. Interestingly, except for pyruvic acid and kynurenine, it was observed that albumin films formed at 2 mg/mL yielded increased amounts of all adsorbed UMs compared to their 0.2 mg/mL counterparts. Not only did the presence of albumin affect pyruvic acid stability, but a thicker film may have more binding sites available for these other metabolites, allowing for their increase in mol% of the total adsorbed species. Bare MNPs showed a complete decrease in adsorbed amount for all types of PBUTs (quinolinic acid > indoxyl sulfate > p-cresol sulfate > hippuric acid > indole acetic acid > kynurenine), while Alb-MNPs showed a moderate to substantial increase of these (except kynurenine) over this period. Most PBUTs have an aromatic ring and an ionic functional group, allowing many potential binding mechanisms, where enthalpy-driven interactions play a role [12,60,61]. Additionally, indoles possess a positive charge that may enable electrostatic interactions with negatively charged Alb-MNPs [61]. Finally, despite partial protein denaturation, Sudlow Sites I and II may be available for binding these metabolites. Ultimately, the high selectivity of Alb-MNPs for PBUTs may be attributed to multiple factors associated with the presence of albumin. Xanthosine and methylhistidine (both molecules are heterocyclic aromatic compounds) have seen a maximum mol% increase in the adsorbed amount for Alb-MNPs. Hence, we can state that Alb-MNPs not only adsorbed protein-bound uremic toxins but also strongly adsorbed non-PBUTs of middle- and low-molecular-weight molecules.

3.4. Uremic Metabolite Binding Characteristics of Albumin vs. Cyclodextrin Films

Previous work has investigated the use of cyclodextrins (CDs) on the adsorption of similar metabolites [30]. The adsorbed nmol of UMs to Alb-MNPs ranged from ~40 to 60 nmol. CD-modified MNPs adsorbed between ~20 and 180 nmols of total UMs for the same mass of particles. These data suggest that the number of binding sites available for metabolites on CD-modified MNPs was significantly larger than that of Alb-MNPs. After 1 h, type D (Film: 25% CD, 75% MPC) exhibited the highest UM (176.45 nmols) adsorption, while type C (Film: 50% CD, 50% MPC) showed the highest (152.58 nmols) adsorption after 4 h. Interestingly, types A (Film: 100% CD) and C demonstrated a significant increase in adsorbed metabolites over time from 16 to 152 nmols, while others showed a decrease from 176 to 35 nmols, though statistically significant changes were observed only for Types A, B (Film: 75% CD: 25% MPC), and D. Similarly, in this study at 1 h, the Alb-MNPs with 2 mg/mL showed the highest (57 nmols) adsorption, while at 4 h, bare MNP was higher (51 nmols). When surfaces had similar zeta potentials, both CD (Type D, −28.3 mV) and albumin (MNPs-APTES-GA-BSA (2), −29.25 mV) coatings showed similar trends, namely, highest adsorbed amounts after 1 h (~60 nmol) and a decrease after 4 h (~42 nmol).
Alb-MNPs yielded two times the concentration of pyruvic acid and increased methylhistidine and creatinine compared to CD-MNPs, where neither were observed. Hippuric acid was concentrated by Alb-MNPs at ~2 to 5 times that of CD-MNPs. It was observed that indole-3-acetic acid did not adsorb to CD films, while its adsorption increased on all Alb-MNPs with time. Likewise, CD-MNPs showed significant spermidine and xanthine, whereas Alb-MNPs adsorbed none. Neither of these surfaces showed the adsorption of alginic acid, asymmetric dimethylarginine, homocysteine, phenolic acid, dimethyl glycine, and spermine.
Based on the heat map comparison presented in Table 4, Alb-modified MNPs have a more comprehensive range of adsorption selectivity from a complex 33 metabolite solution compared to CD-MNPs, which showed only 19 metabolites. While CD-MNPs seemed to have higher adsorption selectivity for only a few compounds (i.e., ethylphenyl sulfate, guanidinopropionic acid, pyruvic acid, and spermidine), Alb-MNPs showed higher adsorption selectivity for a broader range of compounds. These were, in decreasing order, pyruvic acid, methylhistidine, indoxyl sulfate, 4-ethylphenyl sulfate, creatinine, hypoxanthine, quinolinic acid, indole acetic acid, orotic acid, hippuric acid, phenylalanine, uridine, p-hydroxyhippuric acid, trimethylamine N-oxide, uric acid, kynurenine, and guanidinopropionic acid.
Adsorption of 3-deoxyglucosone, hippuric acid, indole-3-acetic acid, indoxyl glucuronide, indoxyl sulfate, p-cresol sulfate, and quinolinic acid on CD- and Alb-modified films are summarized as eluted concentrations (µM) in Figure 9. Even though albumin may have a high binding affinity for PBUTs, the sheer number of binding sites afforded by a CD-rich film yielded increased amounts of adsorbed PBUTs. In most cases, the 100% CD film (particle A, 4 h) had between 2 and 3 times (i.e., 50–126 nmols) the number of metabolites adsorbed by albumin films. This was further reflected by decreased amounts of CDs in the particles labeled B–E, which showed reduced adsorbed amounts of these PBUTs compared to particle A. Interestingly, even pure MPC films (particle E, 4 h) yielded more adsorbed metabolites than albumin films. This suggests that even though albumin was a larger molecule than MPC and MPC-CD polymers, the adsorbed amounts of metabolites were less, likely because most of the albumin structure did not participate in PBUT adsorption.

4. Materials and Methods

4.1. Materials

Chemicals for synthesis: Sodium hydroxide (Fisher Scientific, Waltham, MA, USA), glutaraldehyde (70%, EM, Electron Microscopy Sciences, Hatfield, PA, USA), FeCl2·4H2O, FeCl3·6H2O, NH4OH (25%, Sigma Aldrich, St. Louis, MO, USA), NH2(CH2)3Si(OC2H5)3 (APTES 99%, Sigma Aldrich, St. Louis, MO, USA), BSA (98%, Sigma Aldrich, St. Louis, MO, USA), CH3CH2OH (99.5%, Fisher Scientific, Waltham, MA, USA), and phosphate-buffered saline (PBS) tablets (0.01 M, pH 7.4, HPLC-grade water, filtered at 0.22 µm; Fisher Scientific, Waltham, MA, USA).
Uremic Metabolites: The following chemicals were purchased from Sigma Aldrich: 3-deoxyglucosone (≥75%), 3-indoleacetic acid (≥98%), asymmetric dimethylarginine (≥98%), creatinine (≥98%), L-tyrosine, L-asparagine, dimethylglycine (≥99%), hippuric acid (≥98%), hypoxanthine (≥99%), indoxyl glucuronide (≥98%), indoxyl sulfate, guanidinopropionic acid (≥97.5%), kynurenine (≥98%), methylhistidine, orotic acid (≥98%), p-cresol sulfate (≥95%), phenylalanine (≥98%), pyruvic acid (≥98%), quinolinic acid (≥97%), spermidine (≥99%), spermine (≥97%), trimethylamine N-oxide (≥95%), argininic acid, uric acid (≥99%), uridine (≥99%), xanthine (≥99%), and xanthosine. The 4-ethylphenyl sulfate (98%) was obtained from Apexbio in Houston, TX, USA, and 4-hydroxyhippuric acid (≥98%) from Cayman Chemical, Michigan, MI, USA.

4.2. Synthesis of Iron Oxide Nanoparticles

Iron oxide magnetic nanoparticles were synthesized using the procedure described in detail in [24]. A nitrogen-purged three-necked flask containing FeCl2·4H2O and FeCl3·6H2O in degassed water was stirred at 75 °C. Ammonium hydroxide (25%) was added and stirring increased. After 1.5 h, the nanoparticles were washed, separated using a magnet, and vacuum dried.

4.3. Functionalization with APTES and GA

The functionalization of Fe3O4 nanoparticles (50 mg) was performed as described in [24]. Briefly, the nanoparticles were suspended and sonicated in 5 mL of ethanol, then 35 µL of APTES was added, stirred, and reacted under a nitrogen atmosphere overnight, washed, and suspended in PBS. Following the addition of glutaraldehyde (0.6 mL, 70%), the suspension was stirred for 3 h, washed again, and stored at 4 °C.

4.4. Formation of BSA-Functionalized MNPs

BSA functionalization of nanoparticles was carried out as per the detailed procedures in [24,41]. BSA (0.2 or 2 mg/mL) was dissolved in half the PBS volume, and MNPs (0.08 mg/mL) were sonicated in the other half. The solutions were mixed and shaken at 150 rpm for 2 h. Then, the particles were washed with PBS and stored at 4 °C.

4.5. Particle Characterization

TEM (JEM-ARM200CF S/TEM, JEOL, Houston, TX, USA) at 200 kV was used to determine the particle size and morphology. For TEM, 5 µL of diluted MNPs was placed on a carbon grid and dried overnight. TGA (Pyris 1 TGA, Perkin Elmer, Waltham, MA, USA) was used to assess thermal stability with a 10 °C/min ramp from 25 to 800 °C under nitrogen. Sample weights were 0.84 mg for bare nanoparticles, 1.31 mg for APTES modified, 5.36 mg for MNP-APTES-GA, 0.21 mg for MNP-APTES-GA-BSA (0.2), and 0.42 mg for MNP-APTES-GA-BSA (2). Zeta potential was measured with a Zetasizer (Malvern Instruments, Worcestershire, UK) using 25 µL of 1 mg/mL MNP suspension in 3 mL of DI water. Surface chemistry was analyzed using XPS (Kratos Ultra XPS Spectrometer, Manchester, UK) with 1 mg of freeze-dried sample, as per the previously published protocol [24].
FTIR spectroscopy (Thermo Scientific Nicolet™ iS50 FTIR Spectrometer, Madison, WI, USA) was used to investigate the secondary structure of albumin in both the solution and within the albumin film on the MNP surface. After protein immobilization, 1 mg of freeze-dried powder for all samples was studied. For native BSA, 2 mg/mL of BSA solution was used. The spectra for the solvent and sample were collected at the resolution of 4 cm−1 and 64 scan rates to enhance the signal-to-noise ratio. Each sample spectrum was processed using OMNIC 7.3 software for solvent spectra, baseline correction, and smoothness. The amide-I band (1600–1700 cm−1) characteristic of proteins and the carbonyl stretching were fitted to Gaussian curves, which is the best method to quantify the percentage of secondary structural elements. The protein secondary structure was determined using the Fourier self-deconvolution (FSD) method with OMNIC software. Subsequently, multicomponent peak areas were quantified using Gaussian fitting in Origin 2024 to determine β-sheet, random coil, α-helix, β-turn, and antiparallel β-sheets contents using wavenumber ranges 1614–1634, 1635–1648, 1650–1658, 1662–1678, and 1680–1691 cm−1, respectively.

4.6. Uremic Metabolite Solution

Stock solutions of each UM were made with LC-MS-grade water and combined with uric acid in its powdered form, which was added last. The total volume was adjusted using PBS to form a model uremic toxin solution in 10 mM of PBS at a pH of 7.4.

4.7. MNP Incubation with Uremic Toxin Solution

All tubes were washed with HPLC-grade methanol and vacuum-dried before use. Experimental samples were prepared by adding 0.25 mg of particles to 0.5 mL of UM solution, incubated at 10 rpm at 37 °C for 1 or 4 h, magnetically separated from the metabolite solution, and the solution was removed. These particles were suspended in 100 µL of HPLC-grade methanol and then put in a shaker at 37 °C and 80 rpm for 24 h for the desorption of adsorbed species. MNPs were magnetically separated from the methanol, and methanol samples were centrifuged at 2100 rcf for 20 min to determine whether any particles were still present in the sample. The metabolite–methanol samples were stored at −80 °C until mass spectroscopy could be conducted, and no volume loss was observed.

4.8. Analysis of Adsorbed Metabolites

Plasma samples were analyzed using a targeted MS-based metabolomics approach, combining direct injection (DI) mass spectrometry (MS) and reverse-phase high-performance liquid chromatography (HPLC) tandem mass spectrometry (MS/MS). Chemical derivatization (using 3-nitrophenylhydrazine for organic acids or phenylisothiocynate for amines) was used in this method, followed by extraction, separation, and detection using multiple reaction monitoring (MRM) for metabolite identification and quantification.
Isotope-labeled ISTDs along with other ISTDs were used for accurate metabolite quantification. An Agilent 1290 series HPLC system (Agilent, Palo Alto, CA, USA) and an Agilent reversed-phase Zorbax Eclipse XDB C18 column (3.0 mm × 100 mm, 3.5 μm particle size, 80 Å pore size) were used for online LC-MS/MS with an AB SCIEX QTRAP® 5500 mass spectrometer (AB SCIEX, Redwood City, CA, USA). The controlling software was Analyst® 1.6.3. The mass spectrometer was set to positive/negative electrospray ionization with multiple reaction monitoring (MRM) mode. The raw data for metabolomics were profiled using Multi QuantTM 3.0.3. [62]. Additional details regarding the method, derivatization strategy, separation protocol, MS methods, calibration, and metabolite quantification process were described previously [63].

4.9. Statistical Analysis

Statistical analysis was conducted by quantifying HPLC data of three independent experiments to assess the total adsorbed mols after 1 and 4 h of incubation for all types of nanoparticles (bare, MNP-APTES-GA, and Alb-MNPs). Statistical analysis was conducted using a two-way ANOVA, using Tukey’s multiple comparison tests. Calculated p-values were defined as p < 0.01 and p < 0.001.

5. Conclusions

The adsorption of UMs to albumin-coated MNPs was evaluated using model solutions of thirty-three UMs at concentrations similar to those observed for patients with kidney failure between HD sessions. The albumin solution concentration affected the thickness of the albumin coating formed on the MNPs (Figure 2), where higher solution concentrations (2 mg/mL) formed thicker films and had a lower zeta potential than the lower solution concentration (0.2 mg/mL). TGA and XPS data were used to confirm the higher amount of albumin coating with the increasing albumin solution concentration (Figure 3 and Figure 5). FTIR data were used to evaluate secondary structure changes for BSA and showed that the native structure was relatively retained for 2 mg/mL systems, and some dehydration occurred for thinner films formed at 0.2 mg/mL conditions (Figure 6). It was observed that metabolite binding was not related to solution concentration (Table 4). All types of particles adsorbed pyruvic acid and methylhistidine in substantially higher relative concentrations compared to the original solution concentration. The particles coated with glutaraldehyde and albumin of any concentration showed high selectivity for indoxyl sulfate. Both Alb-MNPs had higher selectivity for ethylphenyl sulphate than bare or glutaraldehyde-coated particles. Additionally, both types of Alb-MNPs showed that as the incubation time increased from 1 to 4 h, there was an increase in the adsorption of individual metabolites (except for pyruvic acid and kynurenine; Table 5). In contrast, bare and glutaraldehyde-coated particles exhibited the opposite trend. Unlike unmodified MNP surfaces, Alb-MNPs adsorbed PBUTs and non-PBUTs of middle and low molecular weight, like xanthosine, methylhistidine, and dimethylglycine. Notably, inconsistent adsorption was reported for indole-3-acetic acid on previously reported CD films, while its adsorption increased on average for all Alb-MNPs with time (Figure 9) [30]. Similar to other work on CD films, adsorption was substantially increased across all surfaces over time for hippuric acid, indoxyl sulfate, and p-cresol sulfate [30]. Moreover, Alb-MNPs exhibited a large amount of adsorption of indole acetic acid, indoxyl glucuronide, 3-deoxyglucosone, quinolinic acid, and kynurenine. CD films were shown to bind more PBUTs than albumin films, likely due to the sheer number of potential binding sites afforded by CD films, but Alb-MNPs showed a greater diversity in adsorbed species. Now that it has been shown that these systems bind UMs from model solutions, further research is needed to evaluate metabolite binding from complex solutions and the general hemocompatibility of these materials.

Author Contributions

I.S., main research work, conceptualization, data curation, formal analysis, investigation, methodology, validation visualization, writing—original draft, writing—review and editing; A.M., research work, experiments; L.Z., research work, data curation; J.Z., research work, data curation; E.L., manuscript—writing and review; D.S.W., conceptualization, writing—review and editing; M.T., reviewing results and manuscript writing; L.D.U., funding acquisition, conceptualization, project administration, resources, supervision, validation visualization, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

L.D.U. acknowledges funding from Natural Sciences and Engineering Research Council: grant number 202102721 and Alberta Innovates: grant number 202102721.

Institutional Review Board Statement

Human plasma experiments followed the guidelines of the research ethics board approvals from Canadian Blood Services 2022-21 and the University of Alberta Pro00002363 and Pro00116764.

Informed Consent Statement

Informed consent statement was obtained from all subjects included in this study.

Data Availability Statement

The data used to support the findings of this study are included in the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PBUTProtein-bound uremic toxin
CKDChronic kidney disease
UMUremic metabolite
HDHemodialysis
MNPMagnetic nanoparticle
p(MPC-co-PMβCD)(Methacryloyloxy)ethyl phosphorylcholine-co-β-cyclodextrin
TGAThermogravimetric analysis
XPSX-ray photoelectron spectroscopy
FTIRFourier-transform infrared spectroscopy
TEMTransmission electron microscopy
FSDFourier self-deconvolution
PBSPhosphate-buffered saline
HPLCHigh-performance liquid chromatography
DIDirect injection
MSMass spectrometry
MS/MSTandem mass spectrometry
MRMMultiple reaction monitoring
BSABovine serum albumin
Alb-MNPAlbumin-modified MNPs
(-NH2-)Amide-I
(-NH-)Amide-II
CDCyclodextrin

References

  1. Jadoul, M.; Aoun, M.; Imani, M.M. The major global burden of chronic kidney disease. Lancet Glob. Health 2024, 12, e342–e343. [Google Scholar] [CrossRef] [PubMed]
  2. Antini, C.; Caixeta, R.; Luciani, S.; Hennis, A.J. Diabetes Mellitus and Cardiovascular Disease: Exploring Epidemiology, Pathophysiology, and Treatment Strategies. Int. J. Epidemiol. 2024, 53, dyad182. [Google Scholar] [CrossRef] [PubMed]
  3. Lim, Y.J.; Sidor, N.A.; Tonial, N.C.; Che, A.; Urquhart, B.L. Uremic Toxins in the Progression of Chronic Kidney Disease and Cardiovascular Disease: Mechanisms and Therapeutic Targets. Toxins 2021, 13, 142. [Google Scholar] [CrossRef] [PubMed]
  4. Ma, Y.-R.; Xin, M.-Y.; Li, K.; Wang, H.; Rao, Z.; Liu, T.-X.; Wu, X.-A. An LC-MS/MS analytical method for the determination of uremic toxins in patients with end-stage renal disease. J. Pharm. Biomed. Anal. 2020, 191, 113551. [Google Scholar] [CrossRef]
  5. Daneshamouz, S.; Eduok, U.; Abdelrasoul, A.; Shoker, A. Protein-bound uremic toxins (PBUTs) in chronic kidney disease (CKD) patients: Production pathway, challenges and recent advances in renal PBUTs clearance. NanoImpact 2021, 21, 100299. [Google Scholar] [CrossRef]
  6. Li, S.; Sharaf, M.G.; Rowe, E.M.; Serrano, K.; Devine, D.V.; Unsworth, L.D. Hemocompatibility of β-Cyclodextrin-Modified (Methacryloyloxy)ethyl Phosphorylcholine Coated Magnetic Nanoparticles. Biomolecules 2023, 13, 1165. [Google Scholar] [CrossRef]
  7. Yang, F.; Liao, M.; Wang, P.; Yang, Z.; Liu, Y. The Cost-Effectiveness of Kidney Replacement Therapy Modalities: A Systematic Review of Full Economic Evaluations. Appl. Health Econ. Health Policy 2021, 19, 163–180. [Google Scholar] [CrossRef]
  8. Klefter, R.; Nielsen, B. Savings in Dialysis Treatment? Artif. Organs 2002, 26, 49–54. [Google Scholar] [CrossRef]
  9. Vanholder, R.; Annemans, L.; Bello, A.K.; Bikbov, B.; Gallego, D.; Gansevoort, R.T.; Lameire, N.; Luyckx, V.A.; Noruisiene, E.; Oostrom, T. Fighting the unbearable lightness of neglecting kidney health: The decade of the kidney. Clin. Kidney J. 2021, 14, 1719–1730. [Google Scholar] [CrossRef]
  10. Vanholder, R.; Snauwaert, E.; Verbeke, F.; Glorieux, G. Future of Uremic Toxin Management. Toxins 2024, 16, 463. [Google Scholar] [CrossRef]
  11. Ma, Y.; Li, S.; Tonelli, M.; Unsworth, L.D. Adsorption-based strategies for removing uremic toxins from blood. Microporous Mesoporous Mater. 2021, 319, 111035. [Google Scholar] [CrossRef]
  12. Gao, C.; Zhang, Q.; Yang, Y.; Li, Y.; Lin, W. Recent trends in therapeutic application of engineered blood purification materials for kidney disease. Biomater. Res. 2022, 26, 5. [Google Scholar] [CrossRef] [PubMed]
  13. Selim, M.M.; El-Safty, S.; Tounsi, A.; Shenashen, M. A review of magnetic nanoparticles used in nanomedicine. APL Mater. 2024, 12, 010601. [Google Scholar] [CrossRef]
  14. Rezaei, B.; Yari, P.; Sanders, S.M.; Wang, H.; Chugh, V.K.; Liang, S.; Mostufa, S.; Xu, K.; Wang, J.P.; Gómez-Pastora, J. Magnetic Nanoparticles: A Review on Synthesis, Characterization, Functionalization, and Biomedical Applications. Small 2024, 20, 2304848. [Google Scholar] [CrossRef]
  15. Salmanpour, Z.; Abolmaali, S.S.; Farahavr, G.; Salmanpour, M.; Tamaddon, A.M. Magnetic nanoparticles and their hybrid biomaterials for drug delivery and theranostic applications in cardiovascular diseases. J. Drug Deliv. Sci. Technol. 2024, 96, 105676. [Google Scholar] [CrossRef]
  16. Wang, L.; Han, K.; Jiang, X.; Mao, C.; Li, X.; Zhou, M. Recent Advances in Blood Toxin Removal Technology. Sustain. Mater. Technol. 2024, 39, e00828. [Google Scholar] [CrossRef]
  17. Li, J.; Han, L.; Xie, J.; Liu, S.; Jia, L. Multi-sites polycyclodextrin adsorbents for removal of protein-bound uremic toxins combining with hemodialysis. Carbohydr. Polym. 2020, 247, 116665. [Google Scholar] [CrossRef]
  18. Niestanak, V.D.; Unsworth, L.D. Detailing Protein-Bound Uremic Toxin Interaction Mechanisms with Human Serum Albumin in the Pursuit of Designing Competitive Binders. Int. J. Mol. Sci. 2023, 24, 7452. [Google Scholar] [CrossRef]
  19. Cunha, R.S.D.; Azevedo, C.A.B.; Falconi, C.A.; Ruiz, F.F.; Liabeuf, S.; Carneiro-Ramos, M.S.; Stinghen, A.E.M. The Interplay between Uremic Toxins and Albumin, Membrane Transporters and Drug Interaction. Toxins 2022, 14, 177. [Google Scholar] [CrossRef]
  20. Li, S.; Tonelli, M.; Unsworth, L.D. Indoxyl and p-cresol sulfate binding with human serum albumin. Colloids Surf. A Physicochem. Eng. Asp. 2022, 635, 128042. [Google Scholar] [CrossRef]
  21. Baki, A.; Remmo, A.; Löwa, N.; Wiekhorst, F.; Bleul, R. Albumin-Coated Single-Core Iron Oxide Nanoparticles for Enhanced Molecular Magnetic Imaging (MRI/MPI). Int. J. Mol. Sci. 2021, 22, 6235. [Google Scholar] [CrossRef] [PubMed]
  22. Al-Abboodi, A.; Alsaady, H.M.; Banoon, S.; Al-Saady, M. Conjugation strategies on functionalized iron oxide nanoparticles as a malaria vaccine delivery system. Bionatura 2021, 3, 2009–2016. [Google Scholar] [CrossRef]
  23. Zhao, L.; Yang, B.; Dai, X.; Wang, X.; Gao, F.; Zhang, X.; Tang, J. Glutaraldehyde Mediated Conjugation of Amino-Coated Magnetic Nanoparticles with Albumin Protein for Nanothermotherapy. J. Nanosci. Nanotechnol. 2010, 10, 7117–7120. [Google Scholar] [CrossRef] [PubMed]
  24. Sharma, I.; Sharaf, M.G.; Pawar, A.; Milley, A.; Unsworth, L.D. Hemocompatibility of Albumin-Modified Magnetic Nanoparticles. Int. J. Mol. Sci. 2024, 25, 11975. [Google Scholar] [CrossRef]
  25. Zhang, M.; Li, L.; Lei, L.; Kang, K.; Xiao, C. Effectively Decontaminating Protein-Bound Uremic Toxins in Human Serum Albumin Using Cationic Metal–Organic Frameworks. ACS Appl. Mater. Interfaces 2022, 14, 55354–55364. [Google Scholar] [CrossRef]
  26. Saito, K.; Fujigaki, S.; Heyes, M.P.; Shibata, K.; Takemura, M.; Fujii, H.; Wada, H.; Noma, A.; Seishima, M. Mechanism of increases in L-kynurenine and quinolinic acid in renal insufficiency. Am. J. Physiol.-Ren. Physiol. 2000, 279, F565–F572. [Google Scholar] [CrossRef]
  27. Vanholder, R.C.; De Smet, R.V.; Ringoir, S. Assessment of Urea and Other Uremic Markers for Quantification of Dialysis Efficacy. Clin. Chem. 1992, 38, 1429–1436. [Google Scholar] [CrossRef]
  28. Vanholder, R.; De Smet, R.; Glorieux, G.; Argilés, A.; Baurmeister, U.; Brunet, P.; Clark, W.; Cohen, G.; De Deyn, P.P.; Deppisch, R. Review on uremic toxins: Classification, concentration, and interindividual variability. Kidney Int. 2003, 63, 1934–1943. [Google Scholar] [CrossRef]
  29. Duranton, F.; Cohen, G.; De Smet, R.; Rodriguez, M.; Jankowski, J.; Vanholder, R.; Argiles, A.; European Uremic Toxin Work Group. Normal and pathologic concentrations of uremic toxins. J. Am. Soc. Nephrol. 2012, 23, 1258–1270. [Google Scholar] [CrossRef]
  30. Li, S.; Sharaf, M.G.; Zhang, L.; Wishart, D.S.; Tonelli, M.; Unsworth, L.D. Adsorption Dynamics of Uremic Toxins to Novel Modified Magnetic Nanoparticles. Macromol. Biosci. 2024, 24, 2300133. [Google Scholar] [CrossRef]
  31. Itoh, Y.; Ezawa, A.; Kikuchi, K.; Tsuruta, Y.; Niwa, T. Protein-bound uremic toxins in hemodialysis patients measured by liquid chromatography/tandem mass spectrometry and their effects on endothelial ROS production. Anal. Bioanal. Chem. 2012, 403, 1841–1850. [Google Scholar] [CrossRef] [PubMed]
  32. Canepa, A.; Filho, J.C.D.; Gutierrez, A.; Carrea, A.; Forsberg, A.M.; Nilsson, E.; Verrina, E.; Perfumo, F.; Bergström, J. Free amino acids in plasma, red blood cells, polymorphonuclear leukocytes. and muscle in normal and uraemic children. Nephrol. Dial. Transplant. 2002, 17, 413–421. [Google Scholar] [CrossRef] [PubMed]
  33. Kopple, J.D.; Swendseid, M.E.; Shinaberger, J.H.; Umezawa, C.Y. The free and bound amino acids removed by hemodialysis. ASAIO J. 1973, 19, 309–313. [Google Scholar] [CrossRef] [PubMed]
  34. Wolfson, M.; Jones, M.R.; Kopple, J.D. Amino acid losses during hemodialysis with infusion of amino acids and glucose. Kidney Int. 1982, 21, 500–506. [Google Scholar] [CrossRef]
  35. Schools, A.; De Vries, P.; Thiemann, R.; Hazejager, W.; Visser, S.; Oe, P. Biochemical and neurophysiological parameters in hemodialyzed patients with chronic renal failure. Clin. Chim. Acta 1989, 185, 91–107. [Google Scholar] [CrossRef]
  36. Landon, J.; Fawcett, J.; Wynn, V. Blood pyruvate concentration measured by a specific method in control subjects. J. Clin. Pathol. 1962, 15, 579–584. [Google Scholar] [CrossRef]
  37. Bain, M.A.; Faull, R.; Fornasini, G.; Milne, R.W.; Evans, A.M. Accumulation of trimethylamine and trimethylamine-N-oxide in end-stage renal disease patients undergoing haemodialysis. Nephrol. Dial. Transplant. 2006, 21, 1300–1304. [Google Scholar] [CrossRef]
  38. Taki, K.; Tsuruta, Y.; Niwa, T. Indoxyl sulfate and atherosclerotic risk factors in hemodialysis patients. Am. J. Nephrol. 2007, 27, 30–35. [Google Scholar] [CrossRef]
  39. Toshimitsu, N.; Naohito, T.; Hideo, Y. RNA metabolism in uremic patients: Accumulation of modified ribonucleosides in uremic serum. Kidney Int. 1998, 53, 1801–1806. [Google Scholar]
  40. Mikhaylova, M.; Kim, D.K.; Berry, C.C.; Zagorodni, A.; Toprak, M.; Curtis, A.S.; Muhammed, M. BSA immobilization on amine-functionalized superparamagnetic iron oxide nanoparticles. Chem. Mater. 2004, 16, 2344–2354. [Google Scholar] [CrossRef]
  41. Rahdar, S.; Rahdar, A.; Ahmadi, S.; Trant, J.F. Adsorption of bovine serum albumin (BSA) by bare magnetite nanoparticles with surface oxidative impurities that prevent aggregation. Can. J. Chem. 2019, 97, 577–583. [Google Scholar] [CrossRef]
  42. Csach, K.; Juríková, A.; Miškuf, J.; Koneracká, M.; Závišová, V.; Kubovčíková, M.; Kopčanský, P. Thermogravimetric Study of the Decomposition of BSA-Coated Magnetic Nanoparticles. Acta Phys. Pol. A 2012, 121, 1293–1295. [Google Scholar] [CrossRef]
  43. Mazario, E.; Forget, A.; Belkahla, H.; Lomas, J.; Decorse, P.; Chevillot-Biraud, A.; Verbeke, P.; Wilhelm, C.; Ammar, S.; Chahine, J.E.H. Functionalization of Iron Oxide Nanoparticles with HSA Protein for Thermal Therapy. IEEE Trans. Magn. 2017, 53, 5200305. [Google Scholar] [CrossRef]
  44. Gonzalez-Moragas, L.; Yu, S.-M.; Carenza, E.; Laromaine, A.; Roig, A. Protective Effects of Bovine Serum Albumin on Superparamagnetic Iron Oxide Nanoparticles Evaluated in the Nematode Caenorhabditis elegans. ACS Biomater. Sci. Eng. 2015, 1, 1129–1138. [Google Scholar] [CrossRef]
  45. Chubarov, A.S. Serum Albumin for Magnetic Nanoparticles Coating. Magnetochemistry 2022, 8, 13. [Google Scholar] [CrossRef]
  46. Hawn, D.D.; DeKoven, B.M. Deconvolution as a correction for photoelectron inelastic energy losses in the core level XPS spectra of iron oxides. Surf. Interface Anal. 1987, 10, 63–74. [Google Scholar] [CrossRef]
  47. Maity, D.; Kale, S.; Kaul-Ghanekar, R.; Xue, J.-M.; Ding, J. Studies of magnetite nanoparticles synthesized by thermal decomposition of iron (III) acetylacetonate in tri(ethylene glycol). J. Magn. Magn. Mater. 2009, 321, 3093–3098. [Google Scholar] [CrossRef]
  48. Park, H.J.; McConnell, J.T.; Boddohi, S.; Kipper, M.J.; Johnson, P.A. Synthesis and characterization of enzyme–magnetic nanoparticle complexes: Effect of size on activity and recovery. Colloids Surf. B Biointerfaces 2011, 83, 198–203. [Google Scholar] [CrossRef]
  49. Chen, X.; Wang, X.; Fang, D. A review on C1s XPS-spectra for some kinds of carbon materials. Fuller. Nanotub. Carbon Nanostructures 2020, 28, 1048–1058. [Google Scholar] [CrossRef]
  50. Sharma, I.; Pattanayek, S.K. Effect of surface energy of solid surfaces on the micro- and macroscopic properties of adsorbed BSA and lysozyme. Biophys. Chem. 2017, 226, 14–22. [Google Scholar] [CrossRef]
  51. Tworek, P.; Rakowski, K.; Szota, M.; Lekka, M.; Jachimska, B. Changes in Secondary Structure and Properties of Bovine Serum Albumin as a Result of Interactions with Gold Surface. Chemphyschem 2024, 25, e202300505. [Google Scholar] [CrossRef] [PubMed]
  52. Givens, B.E.; Xu, Z.; Fiegel, J.; Grassian, V.H. Bovine serum albumin adsorption on SiO2 and TiO2 nanoparticle surfaces at circumneutral and acidic pH: A tale of two nano-bio surface interactions. J. Colloid Interface Sci. 2017, 493, 334–341. [Google Scholar] [CrossRef] [PubMed]
  53. Sharma, I.; Pattanayek, S.K. Effect of characteristics of shear force on secondary structures and viscosity of bovine serum albumin solution. Rheol. Acta 2018, 57, 801–812. [Google Scholar] [CrossRef]
  54. Alhazmi, H.; Albratty, M.; Meraya, A.; Najmi, A.; Alam, M.; Javed, S.; Ahsan, W. Spectroscopic characterization of the interactions of bovine serum albumin with medicinally important metal ions: Platinum (IV), iridium (III) and iron (II). Acta Biochim. Pol. 2021, 68, 99–107. [Google Scholar] [CrossRef]
  55. Shang, L.; Wang, Y.; Jiang, J.; Dong, S. pH-dependent protein conformational changes in albumin:gold nanoparticle bioconjugates: A spectroscopic study. Langmuir 2007, 23, 2714–2721. [Google Scholar] [CrossRef]
  56. Marcelino, A.M.C.; Gierasch, L.M. Roles of beta-turns in protein folding: From peptide models to protein engineering. Biopolymers 2008, 89, 380–391. [Google Scholar] [CrossRef]
  57. Yu, C.; Al-Saadi, A.; Shih, S.-J.; Qiu, L.; Tam, K.; Tsang, S. Immobilization of BSA on Silica-Coated Magnetic Iron Oxide Nanoparticle. J. Phys. Chem. C-J PHYS CHEM C 2008, 113, 537–543. [Google Scholar] [CrossRef]
  58. Miller, T.M.; Wiens, J.P.; Viggiano, A.A.; Ard, S.G.; Shuman, N.S. Thermal Electron Attachment to Pyruvic Acid, Thermal Detachment from the Parent Anion, and the Electron Affinity of Pyruvic Acid. J. Phys. Chem. A 2022, 126, 5545–5551. [Google Scholar] [CrossRef]
  59. Lajtha, A.; Oja, S.; Saransaari, P.; Schousboe, A. Handbook of Neurochemistry and Molecular Neurobiology: Amino Acids and Peptides in the Nervous System; Springer: New York, NY, USA, 2007. [Google Scholar]
  60. Dhondt, A.; Vanholder, R.; Van Biesen, W.; Lameire, N. The removal of uremic toxins. Kidney Int. 2000, 58, S47–S59. [Google Scholar] [CrossRef]
  61. Wei, J.; Li, R.; Zhang, P.; Jin, H.; Zhang, Z.; Li, Y.; Chen, Y. Efficient selective removal of uremic toxin precursor by olefin-linked covalent organic frameworks for nephropathy treatment. Nat. Commun. 2023, 14, 2805. [Google Scholar] [CrossRef]
  62. Zhang, L.; Zheng, J.; Ahmed, R.; Huang, G.; Reid, J.; Mandal, R.; Maksymuik, A.; Sitar, D.S.; Tappia, P.S.; Ramjiawan, B. A High-Performing Plasma Metabolite Panel for Early-Stage Lung Cancer Detection. Cancers 2020, 12, 622. [Google Scholar] [CrossRef]
  63. Zheng, J.; Zhang, L.; Johnson, M.; Mandal, R.; Wishart, D.S. Comprehensive Targeted Metabolomic Assay for Urine Analysis. Anal. Chem. 2020, 92, 10627–10634. [Google Scholar] [CrossRef]
Figure 1. Representative TEM images for (a) bare nanoparticles, (b) APTES modified, (c) GA modified, (d) BSA (0.2) modified, and (e) BSA (2) modified, and SAED images of (f) bare nanoparticles, (g) BSA (0.2) modified, and (h) BSA (2) modified.
Figure 1. Representative TEM images for (a) bare nanoparticles, (b) APTES modified, (c) GA modified, (d) BSA (0.2) modified, and (e) BSA (2) modified, and SAED images of (f) bare nanoparticles, (g) BSA (0.2) modified, and (h) BSA (2) modified.
Ijms 26 05366 g001
Figure 2. Particle size distribution of (a) bare nanoparticles (n = 104), (b) BSA (0.2) modified (n = 46), and (c) BSA (2) modified (n = 53).
Figure 2. Particle size distribution of (a) bare nanoparticles (n = 104), (b) BSA (0.2) modified (n = 46), and (c) BSA (2) modified (n = 53).
Ijms 26 05366 g002
Figure 3. Thermogravimetric analysis for bare, subsequently added APTES, GA, BSA (0.2), and BSA (2) nanoparticles.
Figure 3. Thermogravimetric analysis for bare, subsequently added APTES, GA, BSA (0.2), and BSA (2) nanoparticles.
Ijms 26 05366 g003
Figure 4. ζ-potential for bare and subsequently added APTES, GA, BSA (0.2), and BSA (2) nanoparticles.
Figure 4. ζ-potential for bare and subsequently added APTES, GA, BSA (0.2), and BSA (2) nanoparticles.
Ijms 26 05366 g004
Figure 5. XPS scan for bare (a) and subsequently added APTES (b), GA (c), BSA (0.2) (d), and BSA (2) (e) nanoparticles. (Inset) N 1s, high-resolution scans for APTES- and GA-modified nanoparticles. Accurate fit determined using χ2 and full width at half maximum of component peaks.
Figure 5. XPS scan for bare (a) and subsequently added APTES (b), GA (c), BSA (0.2) (d), and BSA (2) (e) nanoparticles. (Inset) N 1s, high-resolution scans for APTES- and GA-modified nanoparticles. Accurate fit determined using χ2 and full width at half maximum of component peaks.
Ijms 26 05366 g005
Figure 6. FTIR scan of pure BSA, MNPs-APTES-GA-BSA (0.2), and MNPs-APTES-GA-BSA (2).
Figure 6. FTIR scan of pure BSA, MNPs-APTES-GA-BSA (0.2), and MNPs-APTES-GA-BSA (2).
Ijms 26 05366 g006
Figure 7. FTIR Gaussian fitting of amide-I band for (a) pure albumin, (b) MNPs-APTES-GA-BSA (0.2), and (c) MNPs-APTES-GA-BSA (2), respectively.
Figure 7. FTIR Gaussian fitting of amide-I band for (a) pure albumin, (b) MNPs-APTES-GA-BSA (0.2), and (c) MNPs-APTES-GA-BSA (2), respectively.
Ijms 26 05366 g007
Figure 8. Total adsorbed metabolites for all MNP systems after 1 or 4 h of incubation in the model UM solution. Asterisks denote statistical significance between samples, with * representing p < 0.05. Data represent the average ± 1 STD, n = 3.
Figure 8. Total adsorbed metabolites for all MNP systems after 1 or 4 h of incubation in the model UM solution. Asterisks denote statistical significance between samples, with * representing p < 0.05. Data represent the average ± 1 STD, n = 3.
Ijms 26 05366 g008
Figure 9. Quantitative LC/MS analysis of protein-bound uremic toxin (PBUT) adsorption on albumin-functionalized magnetic nanoparticles (Alb-MNPs; left of dashed line) and cyclodextrin-modified magnetic nanoparticles (CD-MNPs; right of dashed line). CD-MNPs (samples A–E) represent increasing molar ratios of cyclodextrin to MPC: A (1:0), B (3:1), C (1:1), D (1:3), and E (0:1). Adsorption profiles were assessed after 1 h and 4 h incubation. Data represent eluted metabolite concentrations (μM), expressed as mean ± standard deviation (n = 3). Blank values indicate concentrations below the detection limit of the method (<LOD).
Figure 9. Quantitative LC/MS analysis of protein-bound uremic toxin (PBUT) adsorption on albumin-functionalized magnetic nanoparticles (Alb-MNPs; left of dashed line) and cyclodextrin-modified magnetic nanoparticles (CD-MNPs; right of dashed line). CD-MNPs (samples A–E) represent increasing molar ratios of cyclodextrin to MPC: A (1:0), B (3:1), C (1:1), D (1:3), and E (0:1). Adsorption profiles were assessed after 1 h and 4 h incubation. Data represent eluted metabolite concentrations (μM), expressed as mean ± standard deviation (n = 3). Blank values indicate concentrations below the detection limit of the method (<LOD).
Ijms 26 05366 g009
Table 1. Composition of uremic toxin solution developed based on literature analysis of the blood of patients with kidney failure.
Table 1. Composition of uremic toxin solution developed based on literature analysis of the blood of patients with kidney failure.
Uremic ToxinPatient
Concentration
(Ave ± 1 SD, mg/L)
Experimental
Concentration
(mg/L)
Solution
Composition
(mol%)
Ref.
3-Deoxyglucosone1.7 ± 1.01.70.131[28]
4-Ethylphenyl sulfate0.242 ± 0.0440.250.015[31]
Argininic acid<0.0770.0770.005[28]
Asymmetric dimethylarginine0.385 ± 0.28840.3850.024[29]
Creatinine136.0 ± 46.013615.023[28]
Dimethyl glycine0.57680.590.071[28]
Guanidinopropionic acid0.288 ± 0.01830.290.028[28]
Hippuric acid247.0 ± 11223616.454[28]
Homocysteine8.1 ± 1.68.10.749[28]
Hypoxanthine2.0 ± 1.620.184[28]
Indole acetic acid2.03 ± 0.382.030.145[29]
Indoxyl glucuronide2.5 ± 0.32.50.101[29]
Indoxyl sulfate53.0 ± 91.5533.106[28]
Kynurenine0.7~1.00.750.045[29]
L-Asparagine7.13 ± 3.77.130.674[32]
L-Tyrosine54.35 ± 16.354.353.747[33]
Methylhistidine0.08459 ± 0.050.0850.006[33]
Orotic acid0.928 ± 0.4890.9280.074[4]
p-Cresol sulfate20.9 ± 12.220.91.387[29]
Phenylacetic acid467.2 ± 10.6467.242.887[29]
Phenylalanine8.92 ± 1.819.250.7[34]
p-Hydroxy hippuric acid4.43 ± 2.794.250.272[35]
p-Hydroxyphenyl acetic acid2.43 ± 2.282.50.205[35]
Putrescine0.00942 ± 0.007590.009420.001[29]
Pyruvic acid11.7 ± 8.611.71.661[36]
Quinolinic acid0.08350.0840.006[26]
Spermidine0.097 ± 0.0450.0960.008[29]
Spermine0.018 ± 0.01620.0180.001[28]
Trimethylamine n-oxide7.49 ± 2.397.51.248[37]
Uric acid83 ± 13836.169[38]
Uridine9.8 ± 11.49.80.501[28]
Xanthine1.5 ± 0.81.50.123[28]
Xanthosine96.6 ± 62.996.64.247[39]
Table 2. XPS chemical composition of bare and subsequently added APTES, GA BSA (0.2), and BSA (2) nanoparticles. All component values are in atomic %.
Table 2. XPS chemical composition of bare and subsequently added APTES, GA BSA (0.2), and BSA (2) nanoparticles. All component values are in atomic %.
SampleMNPsMNPs-APTESMNPs-APTES-GAMNPs-APTES-GA-BSA (0.2)MNPs-APTES-GA-BSA (2)
Fe18.7515.2810.975.941.78
C33.7933.0045.1456.5474.18
O47.4646.7936.8827.2118.11
N-2.013.298.225.47
Si-2.933.712.100.46
Table 3. The percentage of secondary structure content for pure BSA (i.e., in solution) and BSA incorporated into the coating on the MNP surface. Data represents the average ± 2 STD, n = 3.
Table 3. The percentage of secondary structure content for pure BSA (i.e., in solution) and BSA incorporated into the coating on the MNP surface. Data represents the average ± 2 STD, n = 3.
Secondary StructurePure BSAMNPs-APTES-GA-BSA (0.2)MNPs-APTES-GA-BSA (2)
β-sheets/turns10 ± 29 ± 217 ± 2
α-helix65 ± 255 ± 260 ± 2
Random chains5 ± 29 ± 27 ± 2
Extended chains/β-sheets20 ± 221 ± 216 ± 2
Dehydrated β-turns-6 ± 2
Table 4. Heat map illustrating binding specificity *.
Table 4. Heat map illustrating binding specificity *.
Bare MNPsMNP-APTES-GAMNP-APTES-GA-BSAMNP-APTES-GA-BSA
(0.2 mg/mL)(2 mg/mL)
Uremic metabolites1 h4 h1 h4 h1 h4 h1 h4 h
3-Deoxyglucosone0.300.310.320.200.230.240.190.36
4-Ethylphenyl Sulfate1.110.780.991.011.261.240.801.27
Argininic Acid--------
Asymmetric Dimethylarginine--------
Creatinine1.821.061.180.871.191.270.951.18
Dimethyl Glycine0.590.400.500.290.410.590.290.42
Guanidinopropionic Acid0.720.730.670.490.640.820.400.72
Hippuric Acid1.090.880.990.660.841.030.600.85
Homocysteine--------
Hypoxanthine1.340.871.130.731.091.220.941.14
Indole Acetic Acid1.000.840.890.730.981.040.580.87
Indoxyl Glucuronide0.760.530.690.470.630.650.390.60
Indoxyl Sulfate1.070.761.201.461.601.731.011.69
Kynurenine1.801.63-0.210.640.120.940.38
Methylhistidine4.102.443.631.932.022.951.763.20
Orotic Acid1.140.951.080.860.931.040.671.15
p-Cresol Sulfate0.350.270.330.290.310.330.230.34
Phenylacetic Acid--------
Phenylalanine1.140.580.780.530.760.990.520.82
p-Hydroxyhippuric Acid0.810.760.820.600.740.890.500.77
Pyruvic Acid24.1435.1731.5038.7032.8429.0039.7531.73
Quinolinic Acid1.160.771.090.740.991.050.641.02
Spermidine--------
Spermine--------
Trimethylamine N-Oxide0.970.560.730.550.700.830.490.74
Uric Acid0.840.700.830.550.670.820.560.79
Uridine1.010.610.800.590.740.870.520.78
Xanthine--------
Xanthosine0.480.450.410.380.380.410.260.52
* Table 4 displays a heat map of the binding specificity. The values were calculated with the ratio of the normalized, experimentally adsorbed mol% to its mol% in the original solution. Values greater than 1 (dark red) indicate enhanced adsorption relative to the compound’s initial concentration. Adsorption levels were further categorized as follows: 0.99 to 0.8 (medium red), 0.79 to 0.6 (light red), 0.59 to 0.4 (light blue), 0.39 to 0.2 (medium blue), and 0.19 to 0 (dark blue), representing approximately three-quarters, half, one-third, and less than one-fifth of the initial solution concentration, respectively. The “-” symbol denotes cases where calculations were not possible due to detection limits.
Table 5. Percent change in mol% from 1 to 4 h of incubation *.
Table 5. Percent change in mol% from 1 to 4 h of incubation *.
Uremic MetabolitesBare MNPsMNP-APTES-GAMNP-APTES-GA-BSA
(0.2 mg/mL)
MNP-APTES-GA-BSA
(2 mg/mL)
3-Deoxyglucosone1.7−60.05.247.1>30
4-Ethylphenyl Sulfate−41.51.9−1.537.310 to 29
Argininic Acid----−9 to 9
Asymmetric Dimethylarginine----−29 to −10
Creatinine−71.9−36.06.219.2−49 to −30
Dimethyl Glycine−48.5−72.331.030.8<−50
Guanidinopropionic Acid0.5−35.922.444.5
Hippuric Acid−24.2−50.418.930.3
Homocysteine----
Hypoxanthine−53.5−55.210.717.6
Indole Acetic Acid−19.6−21.55.734.0
Indoxyl Glucuronide−44.7−47.23.634.8
Indoxyl Sulfate−42.217.97.340.5
Kynurenine−10.6100.0−425.0−146.1
Methylhistidine−67.6−87.831.644.8
Orotic Acid−20.1−25.410.941.9
p-Cresol Sulfate−30.0−14.76.832.0
Phenylacetic Acid----
Phenylalanine−96.0−46.423.636.3
p-Hydroxyhippuric Acid−7.4−36.816.735.3
Pyruvic Acid31.418.6−13.2−25.3
Quinolinic Acid−51.5−48.55.936.8
Spermidine----
Spermine----
Trimethylamine N-Oxide−73.3−33.216.433.6
Uric Acid−20.3−50.818.929.3
Uridine−65.8−35.014.433.3
Xanthine----
Xanthosine−5.9−7.26.348.8
* Red represents an increase, blue indicates a decrease, and white signifies minimal change.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sharma, I.; Milley, A.; Zhang, L.; Zheng, J.; Lockwood, E.; Wishart, D.S.; Tonelli, M.; Unsworth, L.D. Binding Capacity and Adsorption Stability of Uremic Metabolites to Albumin-Modified Magnetic Nanoparticles. Int. J. Mol. Sci. 2025, 26, 5366. https://doi.org/10.3390/ijms26115366

AMA Style

Sharma I, Milley A, Zhang L, Zheng J, Lockwood E, Wishart DS, Tonelli M, Unsworth LD. Binding Capacity and Adsorption Stability of Uremic Metabolites to Albumin-Modified Magnetic Nanoparticles. International Journal of Molecular Sciences. 2025; 26(11):5366. https://doi.org/10.3390/ijms26115366

Chicago/Turabian Style

Sharma, Indu, Agatha Milley, Lun Zhang, Jiamin Zheng, Ethan Lockwood, David S. Wishart, Marcello Tonelli, and Larry D. Unsworth. 2025. "Binding Capacity and Adsorption Stability of Uremic Metabolites to Albumin-Modified Magnetic Nanoparticles" International Journal of Molecular Sciences 26, no. 11: 5366. https://doi.org/10.3390/ijms26115366

APA Style

Sharma, I., Milley, A., Zhang, L., Zheng, J., Lockwood, E., Wishart, D. S., Tonelli, M., & Unsworth, L. D. (2025). Binding Capacity and Adsorption Stability of Uremic Metabolites to Albumin-Modified Magnetic Nanoparticles. International Journal of Molecular Sciences, 26(11), 5366. https://doi.org/10.3390/ijms26115366

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop