The present disclosure relates to the field of separations, in particular the field of rare earth element (REE) separations.
The present disclosure addresses a new, ecofriendly, bioinspired and energy efficient method for the separation of Rare Earth Elements (REEs) in aqueous feedstocks from mining extraction and on-site processing. REEs are the heavy lanthanide metals in addition to yttrium and scandium. Their unique luminescent, magnetic, and catalytic properties have made them essential to the US economy, with applications ranging from catalysis to luminescent materials in displays, and that include a number of clean energy (rechargeable batteries, wind turbines, solar panels and hybrid vehicles) and other (e.g., satellites, magnets, and guidance systems) technologies.
Although the United States once dominated REE mining and production, environmental restrictions on current separation methods have limited domestic production. The development of robust domestic supply chains is of paramount importance. Accordingly, there is a long-felt need in the field for improved methods of separating and recovering REEs.
This disclosure provides, inter alia, new, green, separation technologies to contribute to the rebuilding of the domestic supply chain. Such technologies can include, e.g., novel peptide-comprising surfactants.
As is known, new separation methods are urgently needed to support an uninterrupted supply of Rare Earth Elements (REEs) used as essential components in advanced technologies important to the national economy. This disclosure provides, e.g., a bioinspired, green REE separation process based on peptide surfactants (PEPS) at air-water interfaces.
PEPS can include a hydrophilic amino acid sequence that selectively binds an REE cation and a hydrophobic sequence that confers surface activity. PEPS added to REE feedstocks can thus associate selectively with REE cations and adsorb to air-water interfaces for recovery in a froth flotation process that obviates the need for solvents as in solvent extraction methods.
The disclosed PEPS design strategy can use lanthanide binding tags (LBTs). LBTs are short peptide sequences that bind an REE and retain that bound state in complex chemical environments in large proteins. The binding loop domain influences the strength and selectivity of REE binding, and can be used to recover a particular REE over another competing REE or even over all competing REEs. As but one example, one can achieve optimal binding of a particular REE, e.g., Tb3+, to exploit its fluorescence properties in protein metallochemistry.
The disclosed technology can be used to, e.g., separate a given REE from one or more other REEs that are comparatively close in size to the given REE. The disclosed technology thus allows for PEPS that can select among REEs by controlling the peptide sequence and peptide selections (of the PEPS) and to enhance the peptides' amphiphilicity to promote their adsorption. This can be enhanced by, inter alia, machine learning-guided mutations, rationally-guided mutations, and by other processes.
The disclosed technology provides peptide constructs that can select among REEs by controlling the constructs' peptide sequence and by tailoring the peptides' amphiphilicity to promote adsorption. The constructs' design can be enhanced by, e.g., machine learning-guided mutations, rationally-guided mutations, and by other processes. Peptides are also amenable to inexpensive, scalable production using genetically engineered bacteria in biotechnological processes. For these reasons, this disclosure utilizes LBTs, as one can make changes to the LBTs' structures to design PEPS.
The presently disclosed technology provides a number of advantageous features, and some of these features are provided below.
Feature 1: PEPS that are surface active in the complexed state and bind REEs over divalent cations, e.g., like those in REE feedstock solutions. One can accomplish this objective by, e.g., addition of hydrophobes to the C-terminus of the LBT sequence. One can also utilize a co-surfactant system to promote synergistic PEPS adsorption. Cosurfactants can be designed to avoid attraction of REE to the interface; to promote the adsorption of the complex; and also to protect the binding loop in the vicinity of the interface. Co-surfactants can also be designed to stabilize foams formed by the disclosed compositions.
Feature 2: PEPS with enhanced selectivity among REEs. The binding loop domain already exhibits selectivity among the larger, lighter lanthanides. One can exploit a strategy based on substitutions at a particular site in the LBT binding domain known to play role in LBT selectivity among REEs. One can also use a parallel synthesis approach to screen libraries for selectivities across the REEs.
Evaluation: One can characterize PEPS:REE complexes at air water interfaces, which characterization can be used to support PEPS design. At the air-water interface, PEPS:REE surface activity can be probed by pendant drop, XPS of sampled monolayers, and by fluorescence confocal microscopy. Microfluidics assays integrated with confocal fluorescence can identify co-surfactant formulations. The molecular composition and arrangement can be probed by X-ray interrogation using X-ray reflectivity, X-ray fluorescence near total reflection, and X-ray absorption spectroscopy. Experiments synergize with molecular simulation, including all-atom simulation to provide 3-D candidate PEPS-cation complex conformation, thermodynamic insights into complex stability, energy landscape for adsorption from bulk to interface. Coarse-grained simulations provide insights into interaction among PEPS-cation complexes on densely-covered interfaces with and without co-surfactants.
Impact: This disclosure presents high scientific and societal impact. The research focuses on detailed molecular understanding of PEPS:REE complexation at the highly anisotropic air-water interface. PEPS have cooperative interactions lacking in simpler surfactants used thus far in ion froth flotation processes, providing a combination of affinity and selectivity that is unattainable in those systems.
The patent or application file contains at least one drawing executed in color. Copies of this patent or patent application publication with color drawing(s) will be provided by the Office upon request and payment of the necessary fee.
In the drawings, which are not necessarily drawn to scale, like numerals may describe similar components in different views. Like numerals having different letter suffixes may represent different instances of similar components. The drawings illustrate generally, by way of example, but not by way of limitation, various aspects discussed in the present document. In the drawings:
The present disclosure may be understood more readily by reference to the following detailed description of desired embodiments and the examples included therein.
Unless otherwise defined, all technical and scientific terms used herein have the same meaning as commonly understood by one of ordinary skill in the art. In case of conflict, the present document, including definitions, will control. Preferred methods and materials are described below, although methods and materials similar or equivalent to those described herein can be used in practice or testing. All publications, patent applications, patents and other references mentioned herein are incorporated by reference in their entirety. The materials, methods, and examples disclosed herein are illustrative only and not intended to be limiting.
The singular forms “a,” “an,” and “the” include plural referents unless the context clearly dictates otherwise.
As used in the specification and in the claims, the term “comprising” can include the embodiments “consisting of” and “consisting essentially of.” The terms “comprise(s),” “include(s),” “having,” “has,” “can,” “contain(s),” and variants thereof, as used herein, are intended to be open-ended transitional phrases, terms, or words that require the presence of the named ingredients/steps and permit the presence of other ingredients/steps. However, such description should be construed as also describing compositions or processes as “consisting of” and “consisting essentially of” the enumerated ingredients/steps, which allows the presence of only the named ingredients/steps, along with any impurities that might result therefrom, and excludes other ingredients/steps.
As used herein, the terms “about” and “at or about” mean that the amount or value in question can be the value designated some other value approximately or about the same. It is generally understood, as used herein, that it is the nominal value indicated ±10% variation unless otherwise indicated or inferred. The term is intended to convey that similar values promote equivalent results or effects recited in the claims. That is, it is understood that amounts, sizes, formulations, parameters, and other quantities and characteristics are not and need not be exact, but can be approximate and/or larger or smaller, as desired, reflecting tolerances, conversion factors, rounding off, measurement error and the like, and other factors known to those of skill in the art. In general, an amount, size, formulation, parameter or other quantity or characteristic is “about” or “approximate” whether or not expressly stated to be such. It is understood that where “about” is used before a quantitative value, the parameter also includes the specific quantitative value itself, unless specifically stated otherwise.
Unless indicated to the contrary, the numerical values should be understood to include numerical values which are the same when reduced to the same number of significant figures and numerical values which differ from the stated value by less than the experimental error of conventional measurement technique of the type described in the present application to determine the value.
All ranges disclosed herein are inclusive of the recited endpoint and independently of the endpoints. The endpoints of the ranges and any values disclosed herein are not limited to the precise range or value; they are sufficiently imprecise to include values approximating these ranges and/or values.
As used herein, approximating language can be applied to modify any quantitative representation that can vary without resulting in a change in the basic function to which it is related. Accordingly, a value modified by a term or terms, such as “about” and “substantially,” may not be limited to the precise value specified, in some cases. In at least some instances, the approximating language can correspond to the precision of an instrument for measuring the value. The modifier “about” should also be considered as disclosing the range defined by the absolute values of the two endpoints. For example, the expression “from about 2 to about 4” also discloses the range “from 2 to 4.” The term “about” can refer to plus or minus 10% of the indicated number. For example, “about 10%” can indicate a range of 9% to 11%, and “about 1” can mean from 0.9-1.1. Other meanings of “about” can be apparent from the context, such as rounding off, so, for example “about 1” can also mean from 0.5 to 1.4. Further, the term “comprising” should be understood as having its open-ended meaning of “including,” but the term also includes the closed meaning of the term “consisting.” For example, a composition that comprises components A and B can be a composition that includes A, B, and other components, but can also be a composition made of A and B only. Any documents cited herein are incorporated by reference in their entireties for any and all purposes.
The present disclosure concerns, inter alia, peptide surfactants (PEPS) that complex selectively to REE cations in a solution. One can exploit their adsorption to rising air bubbles sparged through the solution. The REEs can be collected in the enriched foam formed from these bubbles at the top of the solution. The use of ecofriendly materials contrasts to the current practice of REE separation from the feedstocks into organic solvents, and low energy costs are associated with foam fractionation and gas sparging relative to solvent extraction, which requires multiple liquid-liquid extraction cycles.
One unique feature of this separation are the PEPS, surface active peptides designed to bind selectively with particular REEs. This disclosure describes examples of these molecules and illustrates the molecular details of their complexation with REEs and their adsorption in the complex environment of a fluid interface. PEPS can be based on lanthanide binding tags (LBTs), peptide sequences with hydrophilic residues arranged as a binding loop that forms a coordinating sphere of ligands around the REE cation. The peptide sequences in LBTs can be derived from the calcium binding loops present in natural calcium binding proteins, and LBT candidates can be, e.g., optimized to bind REEs that demonstrate strong fluorescence, terbium and europium. Further hydrophobic peptide sequences can be appended to the loop to make the LBT amphiphilic. By changing the amino acid sequence of the peptide loops, the affinity to a particular lanthanide cation can be tuned for selective separation. These peptides are useable in inexpensive, scalable production using genetically engineered bacteria in biotechnological processes.
Described herein are PEPS-REE complexes, which complexes can be present as aggregates and/or as films. The complexes can be present in a form that is more than one monolayer in thickness.
One can use peptides known to bind to a particular REE to develop PEPS. One can quantify binding and aggregation in water, and will determine PEPS' ability to adsorb and complex with REE cations at the interface. One can also amend the sequence of amino acids in the binding loop to achieve selectivity in REE complexation. Particular positions in the loop can dictate selectivity, and this concept can be explored using the molecular models and simulations developed for the high selectivity LBTs. One can also, as described elsewhere herein, incorporate a comparatively hydrophobic moiety at an end of the PEPS. The hydrophobes can be at either end of the PEPS' linear sequence, with the binding loop between them. When the peptide binding loop wraps the REE, the hydrophobic ends can form a hydrophobic domain. For this reason, one can add hydrophobes to either end of the linear sequence so as to promote their hydrophobicity.
Without being bound to any particular theory or embodiment, at least one other location in the binding loop can contribute to the hydrophobic character of the PEPS. In one example given herein, W (tryptophan) is located at position 7 in the current binding loop. This position can be changed (e.g., in mutants), and one can substitute other moieties (e.g., the non-natural amino acid acridone) at this and other locations. In this way, one can tune the hydrophobicity of the PEPS by selection of the amino acid present in the binding loop.
This disclosure presents both high scientific and high societal impact, as PEPS provide a combination of strong binding and selectivity unattainable in current surfactant REE complexes. Further this disclosure provides methods for REE separation with significant economic impact.
One can apply, for example, combined molecular dynamics (MD) simulation, machine learning (ML) and optimization methods for computational design of LBT peptides with improved lanthanide binding selectivity. A non-limiting computational peptide design scheme can involve the following aspects:
Regular MD simulations were performed for each of the 30 systems including LBT1, LBT1 mutant and LBT3 with 10 different lanthanides. Ten input features include van der Waal radius and Lennard-Jones potential depth of lanthanide, solvent accessible surface area of LBT peptide, the mean distances between the trivalent lanthanide cation and oxygen atoms from the six coordinated residues of the LBT peptide and water obtained from simulations. Metadynamics simulations are then performed for LBT3 with three selected lanthanides to obtain the free energy difference between six different class of metastable binding scenarios to expand the training data set. The total number of input data was ˜50 for ML training. The knock-one-out cross validation was applied during the ML training.
As shown in
The overall effect of mutant of LBT on lanthanides binding can be described by the so called “affinity-selectivity trade-off”. Upon replacing polar residue (N) with negatively charged residue (D) at position 3, binding affinity for light lanthanides (with radius larger than 0.9 Å) increased but the overall binding specificity is decreased. However, upon replacing negatively charged residue (E) with polar residue (Q) at position 12, the LBT mutant is predicted to have an improved binding selectivity against ten different lanthanides but with decrease in overall binding affinity.
Solvent extraction is widely used for the separation of rare earth element trivalent cations (REEs or Ln3+) from an aqueous phase into an organic solvent. This process is energy intensive and environmentally unfriendly, requiring large volume of organic solvents and organo-phosphate surfactants. Here, we exploit the high affinity of a surface-active Lanthanide Binding Tag (LBT) peptide (LBT1, YIDTNNDGWYEGDELLA), that coordinates selectively with Ln3+ ions for its use in bioinspired/eco-friendly extraction processes in which the complexed LBT/Ln3+ peptide adsorbs to the air/aqueous interfaces of bubbles for foam recovery. To understand the surface activity and identify the bound cationic species at the air-aqueous interface, we characterized the surface molecular adsorption and arrangement of LBT1, the more surface active LBT1-LLA (YIDTNNDGWYEGDELLALLA), and the less negatively charged LBT1-3 (YIDTNNDGWYEGNELLA).
X-ray reflectivity (XR) and x-ray fluorescence near total reflection (XFNTR) measurements of the adsorbed layer were used to compute the surface concentration of the peptide and the Ln3+ cation, and electron density profile (EDP) of the interfacial layers.
The addition of three hydrophobic residues to LBT1 increased the adsorption of Tb3+ ions to the air-aqueous interface. We demonstrated that this cation adsorption enhancement was also promoted by electrostatic interactions between the charged LBT1-LLA/Tb3+ complexes and free trivalent ions in solution. Moreover, that bridging of complexes via O—Tb—O association can be eliminated by substituting negatively charged groups that do not participate in the selective coordination with neutral amino acids. LBT peptides can promote the adsorption of lanthanides selectively and this affinity can be controlled by varying the concentration of ions in solution. The ability to tune the amino acid sequence of these surface-active molecules to either improve their adsorption and/or selectivity with REEs allows for an advantageous, green, eco-friendly, and selective separation method of REEs.
Rare earth elements (REEs) are those elements between Lanthanum and Lutetium on the periodic table, also known as Lanthanides. Yttrium and Scandium are often included in this category of elements as they shared chemical and physical similarities with the group and exist in nature with other Lanthanides (Ln3+).
The unique properties of REEs make possible their use in a wide range of industrial applications, such as electronics, catalysis, clean energy, batteries, magnetics, and others. Deposits of these elements are found all over the world in the Earth's crust. The major ore deposits containing large amount of these elements are in China, the United States, Australia, India, and Russia.
Although the United States possesses domestic resources of REEs, the country is currently totally reliant on imports. In fact, in 2020 all the REEs consumed by the United States were imported, principally from China, Estonia, Japan, and Malaysia. The low mineable concentration of REE in ores makes it economically infeasible to use current technology for their extraction in the US. Moreover, environmental regulations imposed by the US government in the 1990s have promoted the reduction in local mining of these precious metals. While locating new geological deposits might increase the domestic supply of REEs, the development of new eco-friendly extraction and recycling technologies will certainly increase the local production of these elements by permitting the use of existing sources of REEs.
Before the separation of REEs from a concentrated solution, ores go through a series of physical separation and hydrometallurgical processing. Deposits or ores containing REEs are distinctive from each other. Therefore, the technologies and mineral processing implemented for obtaining concentrated solutions of REEs are broad and depend on the type of ore or mineral that is being processed. Ore beneficiation, mineral concentrate decomposition, and rare earth leaching are techniques that are commonly used for the preconcentrating of REEs solutions. Likewise, several chemical separation techniques can be used for the separation of REEs from concentrated solutions containing these elements, such as ion exchange, chromatography, solvent extraction processes, and others. The solvent extraction process takes advantage of the ability of metal ions to be transported across the interface between an aqueous solution and a non-miscible organic solution.
Generally, the aqueous solution contains the valued ions and soluble impurities, while the organic solution contains an extractant. The desired solute is initially dissolved in the aqueous solution, but eventually is distributed between the two phases until equilibrium is achieved. The time of contact or mixing of the two solutions depends on the affinity of the solutes for the two solvents, and after the two phases are separated the liquid loaded with the valuable solute is called “extract”, while the liquid containing only undesirable solutes and impurities is called “raffinate”.
Separation and purification of REEs have relied upon solvent extraction because of the extensive advantages of the process, such as simple fast and continuous operation, mild process conditions, and inexpensive handling of large quantities of materials. Solvent extraction process for the separation of REEs requires an abundance of solvents. Due to the high viscosity of the extractants use for the collection or REEs, solvents like kerosene and certain aromatics are used to dissolve these extracting components.
Although the different techniques employed throughout the REEs separation and purification process—from mining to solvent extraction—have the potential for environmental damage when not controlled and managed properly, the environment impact is mainly inherited from the solvent extraction stage.
For REEs separation and purification from bastnaesite, solvent extraction contributes approximately 75% to the terrestrial acidification; more than 60% to the global warming potential; more that 50% to the terrestrial, freshwater, and marine eutrophication; and around 70% to the resource depletion for water on the entire REEs extraction process that is represented by mining, beneficiation, acid roasting, leaching, and solvent extraction.
Ion foam flotation (IFF) is a promising eco-friendly approach for the extraction and purification of REEs. IFF is a separation technique that involves adding soluble surfactants or extractants with opposite charges to those of the target ions into solution containing the desired elements to form an extractant-ion complex that will be collected by passing gas bubbles throughout the aqueous solution. To date, IFF has been mainly applied for adsorption techniques, industrial wastewater treatment, and recovery of precious metals. Recently, IFF has received attention for its application in the separation and purification of REEs due to its simplicity, low energy requirements, rapid operation, and eco-friendly approach. However, the use of this technique for the separation and purification of REEs has been considered as an intrinsically immature technology and has not yet found applications at a large commercial scale.
IFF studies have investigated selective separation of REEs. They have demonstrated that the affinity for REEs is higher for higher atomic number elements over those with a smaller number. as well as a significant influence in the selectivity of REEs by the pH of the media and the presence of chloride ions in solution. For the surfactants used in these studies, the electrostatic attraction/coordination between the REEs and the extractants determine the selectivity towards the different trivalent cations, which poses an enormous challenge for simple surfactants with non-cooperative binding interactions.
Calcium-containing proteins have been explored to study their coordination with lanthanides ions based on the similarity of trivalent Ln3+ to divalent Ca2+ in ionic radii and metal coordination chemistry. Furthermore, researchers have engineered biomolecules that bind naturally with Ca2+ to enhance the affinity to Ln3+ ions and in some cases to increase the preference towards specific elements.
Lanthanide Binding Tags (LBTs) are short peptides (constituted by 15 to 20 amino acids) designed and engineered based on EF-hand motifs of calcium-binding proteins. LBT1, with the amino acid sequence YIDTNNDGWYEGDELLA, is a linear peptide of 17 amino acids that has been optimized by screening methods to coordinate selectively with Ln3+ ions. This peptide possesses a higher affinity for Tb3+ compared to the rest of the trivalent lanthanides. LBT1 is composed of hydrophilic amino acids sequences that selectively bind Ln3+ ions, and hydrophobic amino acids located at the end of the C-terminus, which form a hydrophobic cluster that is considered to increase the Ln3+ selectivity.
In this discussion, we explore the affinity of this peptide with Tb3+ ions as well as the hydrophobicity of the molecule for a bio-inspired, green REEs separation process at air-aqueous interface. Furthermore, we enhanced the hydrophobicity of LBT1 by extending the C terminus of the peptide with the addition of three hydrophobic residues (LLA) to improve the adsorption and extraction of REEs. Without being bound to any particular theory, one can hypothesize that the addition of these three residues to the backbone of the peptide will not affect its coordination and selectivity with Ln3+ ions. Likewise, we expect that the hydrophobic residues attached to the C-terminus of both LBT1 and LBT1-LLA, which do not participate in the coordination loop will promote the adsorption of the peptides to the air-aqueous interface without compromising the cation complexation.
The disclosed separation process can include several steps. Initially, peptides added to the ionic feedstocks coordinate selectively with REEs cations. Next, adsorption of complexes containing REEs to air-aqueous interfaces of bubbles will occur. This process can be initiated by introducing gas bubbles through the feedstock. Finally, LBTs-REEs complexes will “float” to the top of the solution where they will be collected as a foam for REEs recovery.
This disclosure reviews the fundamentals on the adsorption of LBT peptides in the absence and presence of Ln3+ cations to address the second step in the proposed separation process. The surface activity of LBTs:REEs complexes is probed by pendant drop tensiometry (PDT) to confirm the adsorption of species to the air-aqueous interface. Detailed characterization of the interface molecular composition and arrangement is determined by X-ray interrogation of adsorbed complexes using X-ray reflectivity (XR), and X-ray fluorescence near total reflection (XFNTR).
LBT1: YIDTNNDGWYEGDELLA (purity 98%) and LBT1-LLA: YIDTNNDGWYEGDELLALLA (purity ≥95%) both labeled at the N-terminus with a free amine and labeled at the C-terminus with a free acid, and LBT1-3: YIDTNNDGWYEGNELLA (purity ≥95%) labeled at the N-terminus and at the C-terminus with a free amine, were purchased from GenScript (Piscataway, NJ, USA), diluted to a stock concentration of 150 μM in buffer solutions containing 100 mM of NaCl (purity ≥99.5%, Sigma-Aldrich) and 50 mM of MES (purity ≥99.5%, Sigma-Aldrich) at a pH of 6, and used without additional purification. Anhydrous TbCl3 (purity ≥99%) was purchased from Sigma-Aldrich and diluted to a stock concentration of 25 mM in the same buffer solution containing 100 mM of NaCl and 50 mM of MES. Buffer solution is filtered using a 0.22 μm polytetrafluoroethylene filters. Ultrapure water is obtained from a Milli-Q water filtration unit (EMD Millipore) with a resistivity of 18.2 mΩ cm and used for preparation of buffer solution.
Fluorescence resonance energy transfer (FRET) between Tryptophan (Trp) and Tb3+ was monitored with a Tecan Spark plate reader, using a 96 well plate in triplicate. Trp was excited at 280 nm and emission spectra was detected at 545 nm. As the concentration of Tb3+ increases in solution for a constant concentration of peptide, the collected emission became more intense until the luminescence intensity became constant (maximum value for saturation of peptide).
The collected data was fitted to produce thermodynamic parameters for the association of the peptide with Tb3+. To obtain thermodynamic association parameters for the rest of the series of lanthanides, displacement assays were carried out. Tb3+ cations were added to the peptide solution, followed by increasing amounts of other Ln3+. As more of the other Ln3+ was added, Tb3+ was displaced. The other lanthanides do not emit at 545 nm, so this displacement promoted a reduction in detected luminescence, that allowed to determine binding free energy and affinity constants for the rest of the Ln3+ series one at the time.
Dynamic surface tension was measured using a pendant drop tensiometer (Attension Theta, Biolin Scientific, Stockholm, Sweden). For this we used a drop profile of the different solutions of LBTs or LBTs:Tb3+ at different concentrations to indirectly measure surface tension. Drops with volume of 14 μL were formed and suspended from a 16-gauge metal needle attached to a volume control system. For each run drop images were captured over time until an approximately constant surface tension was achieved. The surface tension values were computed by fitting the experimental drop shape to a theoretical profile governed by the Laplace equation of capillarity.
X-ray reflectivity (XR) and x-ray fluorescence near total reflection (XFNTR) measurements were conducted at NSF's ChemMatCARs, sector 15-ID-C experimental hutch at the Advanced Photon Source in Argonne National Laboratory (Lemont, IL) using 10 keV x-rays.
The liquid sample was placed in a trough of maximum surface area of 21 cm2 within the x-ray reflectivity apparatus at 15-ID-C. The trough was filled with approximately 6 mL of solutions of LBTs or LBTs:Tb3+. The trough was contained in a sealed aluminum box, resting on a vibration isolation table. After the trough was filled, the box was closed. A slight overpressure of helium was kept in the box to reduce x-ray background scattering. All measurements were taken after approximately 2 hours of the filling process and at room temperature.
X-ray reflectivity measures the interfacial electron density profile. X-ray reflectivity data were measured as a function of wave vector transfer QZ=(4π/λ)sin (α) along the surface normal to cover the range 0.016 Å−1<QZ<0.6 Å−1, where a is the angle of incidence and A is the x-ray wavelength λ=1.24 Å. XR data was measured with a Pilatus 200K area detector. Reflectivity was measured multiple times to check for stability. XR measurements were fit to a model functional form to compute the electron density ρ(z) along the z direction perpendicular to the interface, and averaged over the x-y plane of the surface, with the model function represented by a sum of error functions:
XFNTR measures the surface number density of fluorescing metal ions (number per Δ2). X-ray fluorescence intensity measurements were measured for a range of QZ near the critical value Qc for total reflection. The x-ray fluorescence spectrum was measured by a Vortex-90-EX silicon drift detector placed above the interface and spectra were normalized to the incident beam intensity. The XFNTR signal was obtained by integrating the fluorescence peak of Tb3+ over the acceptance volume in the aqueous phase.
MD simulations are performed to modeling the LBT-Tb3+ binding complex in the aqueous solution using GROMACS package. Peptides were modelled using CHARMM36 force field. Terbium cation was modelled using the modified CHARMM force field, which were designed to match the hydration structure and hydrogen free energy from experimental measurements. Solvent was modelled using the modified Tip3p water model under neutral pH condition. Unless differently stated, we use periodic boundary conditions in the x-, y-, and z-directions. Sodium and chloride ions were used to neutralize the system and the concentration of NaCl was 0.1 M, which is comparable to the experimental conditions. Particle Mesh Ewald algorithm was adopted for the calculation of long-range electrostatic interactions. The integration time step was set to 2.0 fs, and the LINCS algorithm was employed to constrain the lengths of all chemical bonds involving hydrogen atoms at their equilibrium values. The solvated system was first energy minimized using the Steepest Descent (SD), while algorithms were used to remove unfavorable contacts. The isochoric isothermal (NVT) simulation were then performed at room temperature of 298K using a stochastic velocity rescaling algorithm for 5 ns. After the equilibration stage, isobaric isothermal (NPT) simulation were performed under room temperature and ambient pressure (1 bar) a using velocity rescaling thermostat and Parrinello-Rahman barostat.
Well-tempered metadynamics simulations were performed to construct free energy landscape of the LBT1/Tb3+, LBT1-LLA/Tb3+, and LBT1-3/Tb3+ binding complex using the GROMACS package 2020.2 and Plumed 2.6.2 version. The initial configuration of the LBT1/Tb3+ binding complex was obtained from the X-ray measurements (PDB code: 1TJB). The initial configuration of LBT1-3/Tb3+ binding complex was obtained by residue mutation using the Scwrl4 program. The initial configuration of LBT1-LLA/Tb3+ binding complex was obtained by extending the extra -LLA motif from C-terminal of LBT1 peptide with extended conformation using Discovery Studio. The buffering distance of LBT1-3/Tb3+ complex was set to be 1.2 nm which results in a cubic box with length around 5 nm.
Metadynamics simulations of three systems were performed after using the final configurations obtained from a short NPT simulation. Here, we used two collective variables (CV). First, the coordination number (CN) of Tb3+ which refers to the number of oxygen atoms from the LBT peptide that have a separation distance with Tb3+ less than 0.27 nm which, do not include the oxygen of water. Second, the root mean square deviation of α carbon (Cu-RMSD) of the 1st-12th residues (corresponding to position 1 to 12) with respect to the reference structure, (PDB code: 1TJB). The height of the gaussian potential was 1.0 kJ/mol which was deposited every 500 steps (PACE). The widths (SIGMA) of gaussian potentials for CN and Ca-RMSD were 0.3 and 0.05, respectively. The bias factor for well-tempered metadynamics was set to be 3. In addition, the upper wall restraining potential was imposed at Cα-RMSD=0.6 nm and the lower wall restraining potential was imposed at CN=5. Both upper and lower wall potential used a force constant (KAPPA) of 300 kJ and a power (EXP) of 2. The grid boundaries (GRID_MIN and GRID MAX) for CN and Cu-RMSD were set to be 2-11 and 0-1.1 nm, respectively. The number of bins for every collective variable (GRID BIN) was set to be 500. Eight replicas with different random seeds were run in parallel with each replicate lasting around 2 μs.
MD simulations at the air-aqueous interface of three short peptides known as lanthanide binding tags (LBT) were performed. The first peptide referred to as LBT1, is modeled using the structure 1TJB from the protein data bank, which consists of a complex formed by the peptide and a trivalent lanthanide cation. The second peptide LBT1-LLA is built by taking the structure of LBT1 and adding two leucine (L) amino acids and one alanine (A) amino acid. At the two ends, both peptides are terminated by the charged groups NH+3 and COO−. The net charge of these two peptides is −5e. The third peptide LBT1-3 is built by taking the structure of LBT1 and mutating the aspartic acid (D) in position 11 by asparagine (N) such that D11 becomes N11. At the two ends, the peptide is terminated by the charged groups NH+3 and CONH2. The net charge of this peptide is −3e which is exactly compensated when it forms a complex with trivalent ions.
The configurations of the peptide molecules employed in the microsecond MD simulations are consistent with the well-tempered metadynamics simulations. To simulate the molecules in bulk, we employed a cubic simulation box of side L=8 nm. The simulation box contains one peptide molecule forming a complex with the ion Tb+3, 33 Na+ ions, 31 Cl− ions, and 16780 water molecules, equivalent to a 0.1 M of NaCl. We used the same composition to simulate the complexes near an air-aqueous interface by extending the simulation box in the z-direction, namely, the simulation box dimensions are 8 nm, 8 nm, and 32 nm, in the x, y, and z directions respectively. We used periodic boundary conditions in the three directions. The systems containing a single molecule are simulated for at least 2 microseconds. To equilibrate the molecules in bulk and at the interface we performed microsecond MD simulations using a time-step of 2.5 fs at T=298 K. The bulk systems are simulated using the NPT ensemble while for the systems with an interface we use the NVT ensemble. The temperature is controlled using the Nose-Hoover thermostat, and the pressure is maintained using the Parrinello-Rahman barostat.
The adsorption of Tb3+ is simulated employing a simulation box of 10 nm by 10 nm by 40 nm, in the x, y, and z directions, respectively. The aqueous phase occupies about 16 nm in the z-direction of the simulation box, and we use periodic boundary conditions in the three directions. At t=0, the peptide-Tb3+ complexes are placed near the air-aqueous interface. The simulations are performed for one microsecond. Equilibration of the adsorption of cations is achieved at about 600 ns of the simulation. We use the last 400 ns of the simulation trajectory to calculate the average ions adsorption at the interface.
Thermodynamic Study of LBTs Peptide with Lanthanide Cations
Competitive titration measurements were taken for all peptides LBT1, LBT1-LLA, and LBT1-3 to evaluate the binding affinity of peptides with a series of Ln3+ cations. The binding free energy values, shown in
Although measured affinity constants for LBT1 reported in the inset table of
Structures of LBT-Tb3+ complexes. Conventional MD simulations are limited in sampling the rugged conformational free energy landscape of the LBTs-Ln3+ binding complexes. Well-tempered metadynamics simulations add bias potential to the system total potential along two collective variables to enhance sampling of broad conformational space within reasonable time scale.
These simulation findings are consistent with the binding energy measurements show in
Final conformations of the microsecond MD simulations for all the LBT/Tb3+ complexes are shown as inset snapshoots in
Quasi-equilibrium surface tension measurements of solutions containing LBT1, LBT1-LLA, or LBT1-3 for bulk concentrations of 25 μM, 50 μM, and 100 μM in the absence and presence of Tb3+ cations, added in the form of TbCl3 salt are shown in
Surface tension values for the three peptides in the presence of Tb3+ cations at different bulk concentration and a ratio peptide:Tb3+ of 1:4, are displayed in
Microsecond MD simulations of complexes at the air-aqueous interface indicate configurational differences between LBTs-Tb3+ complexes in bulk and at the interface. In the LBT1 complex in bulk and at the interface, the alpha-helix appears and disappears during the microsecond MD simulations whereas in the LBT1-LLA and LBT1-3 bound peptides the alpha-helix persists. These conformational changes affect the coordination of the trivalent ion in the complex. The coordination distances between the Tb3+ ion and the residues are reported in Table S2. In general, either residues D1 or D5 are coordinated only through a single oxygen (monodentate) but the other is coordinated through both oxygens (bidentate). In all the molecules in bulk and at the interface, the residues E9 and E12 are coordinated through both oxygen (bidentate). The coordination with N3 and W7 is unstable. In LBT1 in bulk and at the interface both N3 and W7 detach, and well as in LBT1-LLA at the interface (see
X-ray reflectivity is a technique that uses the x-ray reflection curves resulting from grazing incident x-ray beam to probe the electron density variation ρ(z) normal to the surface z but averaged over the x-y plane. Interfacial molecular quantification is then computed by correlating the measured electron density and the known electron density of the chemical species present at the interface. X-ray reflectivity measurements from the air-aqueous interface and electron density profiles of LBT1/Tb3+, LBT1-LLA/Tb3+, and LBT1-3/Tb3+ adsorbed to an air-aqueous surface are shown in
The electron density profile (EDP) of each peptide-cation at different concentrations shown in
XR measurements and EDPs of adsorbed layers at the air-aqueous interface from solutions containing LBT1, LBT1-LLA, or LBT1-3 are shown in
Surface concentration of all peptides obtained from the XR and XFNTR analysis shown in
The recruitment of trivalent cations to the air-aqueous interface is caused by the adsorption of LBTs/Tb3+ complexes from bulk solution to the air-aqueous surface. XFNTR was used to determine the interfacial density of trivalent cations.
The interfacial density of Tb3+ cations as a function of the trivalent salt subphase concentration, illustrated in
To monitor the location of the hydrophobic section of the molecule with respect to the interface, we investigate the average density profile of different sections of the molecule setting the dividing surface (z=0) at the plane at which the water density is one-half of the bulk density. The most hydrophobic section of the molecules is formed by the amino acids in positions 13 to 15 in LBT1 and LBT1-3, and in positions 13 to 18 in LBT1-LLA (see Table S1). The density profiles in
The change of the free energy (DOG) between the molecules in a bulk solution and at the air-aqueous interface, shown in
The density profiles of the complexed molecules represented in
Quantification of surface concentration of LBT peptides and Tb3+ cations allow to determine the number of cations per peptide at the interfacial zone (ΓTb
Although a ratio ΓTb
XR and XFNTR measurements of adsorbed layers from solutions containing LBT1-LLA at a constant concentration of 100 μM and Tb3+ cations at concentrations varying from 25 μM to 100 μM were taken to determine whether the adsorption of excess cations is induced by the excess unbound cations present in solution.
Without being bound to any particular theory or embodiment, the secondary coordination observed for LBT1 and LBT1-LLA—that promotes the network formation may take place outside of the binding loop of the LBT peptide—may be independent of the ion size, hydration number, or acidity of the metal. Thus, this binding is not discriminatory between different Ln3+ cations and hence, not desire in a selective REEs separation process. LBT1-3 peptide, a C-terminal amidated peptide with a total net charge of −3 in the apo-state, does not promote a secondary binding according to the results represented in
One-microsecond MD simulations are performed to equilibrate the interfacial systems containing the peptide molecules at different surface area concentrations.
Finally, in the secondary coordination for LBT1, the Tb3+ ions coordinate preferentially with carboxylate groups (COO—) from the D11 amino acid, whereas in LBT1-LAA, the Tb3+ ions have higher coordination with the carboxylate groups from the C-terminus than with the D11 amino acid (
Competitive adsorption between Tb3+ and La3+ when using LBT peptides as carriers to air-aqueous interfaces was studied. These cations were selected for the competition assays because of the greater difference in affinity with the peptides (see
Surface densities of LBT1-LLA, Tb3+, and La3+ cations were calculated and are represented in
Lastly, to evaluate the affinity of LBT1-LLA at the air-aqueous interface, interfacial Tb3+/La3+ ratios were calculated and are represented in
Analogously to LBT1-LLA, surface densities of LBT1-3, Tb3+, and La3+ cations were calculated for a fixed peptide bulk concentration of 100 μM and different equimolar concentration of lanthanides in solution (see
Surface concentrations of Tb3+ and La3+ from solutions containing equimolar bulk concentrations of ions and LBT1-3 at a bulk concentration of 100 μM are reported in
Interfacial Tb3+/L3+ ratios for different Ln3+ bulk concentrations were also calculated and are shown in
The changes in affinity observed for the neutrally charged-complexed LBT1-3, indicate that changes in affinity of LBT peptides is not only caused by secondary associations occurring outside of the binding loop. The modifications on the affinity of LBT peptides studied here with lanthanides cations might be caused by changes in conformations of complexes once they adsorb to that air-aqueous interfaces (see snapshots in
These modifications on selectivity seem as well to be dependent on the bulk [Ln]0/[LBT]0. For concentration of peptide of 200 μM ([Ln]0/[LBT]0=2) and lower, there is a greater affinity for La3+ over Tb3+.
Moreover, interfacial Ln3+/LBT1-3 ratios are lower than 1 (see
As the lanthanides bulk concentration increases (higher than 200 μM), a higher affinity towards Tb3+ is observed, with the surface concentration of the two ions approximately the same for both peptides at [Ln]0/[LBT]0 equal to 3, and a greater affinity for Tb3+ over La3+. While this greater affinity for Tb3+ might be related to the higher concentrations of lanthanides in solution, a fully saturated interface is observed for LBT1-LLA, suggesting a rise in molecular interactions between interfacial complexes. Simultaneously, the number of Ln3+ per LBT1-3 is higher for this ionic concentration regime, which indicates the presence of mostly complexes at the interfacial zone that results in different kind of interfacial molecular interactions (not negatively charges molecules are present at the interface).
Finally, for Ln3+ bulk concentration higher than 300 μM, a higher number of ions per peptide is observed for LBT1-LLA (see
In this discussion, we showed that Lanthanide Binding Tag peptides promote the adsorption of Ln3+ cations to air-aqueous interfaces by coordinating with cations in bulk solution and subsequently diffusing to the interface. Moreover, we showed that while the LBT-Ln3+ complexes experience conformational changes once they adsorb to the interface, their binding pockets formed in solution are stable at the interfacial zone. Furthermore, the more hydrophobic peptide (LBT1-LLA) can improve the adsorption of Tb3+ ions to the air-aqueous interface by up to 30% due to its higher surface activity. This enhanced adsorption is promoted not just by the additional hydrophobic residues of the peptide, but as well by the capacity of the peptide to electrostatically bridge individual complexes with extra free ions in solution. Undesirable association of ions with negatively charged groups from the side chain of the peptide can be eradicated by eliminating the charges participating in this secondary coordination without significantly changing the relative affinity of the peptide for the trivalent cations. LBT peptides studied in this work showed to be selective along the Lanthanide series at air-aqueous interfaces, with affinities different to the ones observed in bulk solution. These affinities indicated to be dependent in the bulk Ln/LBT concentration ratio. Moreover, neutral LBT/Ln3+ complexes, such as LBT1-3/Ln3+ exhibited a greater selectivity at the air-water interface when comparing it with the bulk one. Finally, the ability to modify the amino acid sequence of LBT peptides and the concentrations of species in solution for controlling the selectivity and extraction of Lanthanide cations to air-aqueous interfaces makes possible their use in green ion foam flotation separation processes of REEs.
Free energy for different basins obtained from well-tempered metadynamics simulations. Convergence of the enhanced sampling in the well-tempered metadynamics simulations was checked by calculating free energy values at different simulations time, reported in FIG. S2Y and calculated using the following equation:
Where Fbasin is the free energy value integrated over a given two-dimensional area covered by collective variables s1 and s2 which corresponds to the coordination number (CN) of Tb3+ and the root mean square deviation of α carbon (Cα-RMSD) of the binding loop domain of LBT peptide respectively. kB is the Boltzmann constant, T is the temperature.
The first column is the name of the peptide. NT and CT are the N-terminus and C-terminus respectively with charges corresponding to the experimental studies (pH 6). The last row is the position index for each residue. The amino acids coordinating with the ion at t=0 in simulations are colored in blue. The residues that participate in the alpha-helix structure are highlighted using boldface italics.
dO1D1 and dO2D1 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the aspartic acid D1; dO1D5 and dO2D5 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the aspartic acid D5; dO1E9 and dO2E9 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the glutamic acid E9; and dO1E12 and dO2E12 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the glutamic acid E12.
dO1D1 and dO2D1 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the aspartic acid D1; dO1D5 and dO2D5 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the aspartic acid D5; dO1E9 and dO2E9 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the glutamic acid E9; and dO1E12 and dO2E12 are the distances from the Tb3+ cation to the carboxyl oxygen atoms from the glutamic acid E12.
Free energy calculations from MD simulations. The change of the free energy is calculated using the slow-growth method in thermodynamic integration to transform the system from state A to state B as follows:
l is a coupling parameter that varies from 0 to 1 to modulate the interaction between the target molecule and the medium that is gradually turned in the system's Hamiltonian, such that the interaction between the medium and the molecule is switched off at state A (l=0) and is on in state B (l=0).
X-ray Reflectivity (XR) Data Analysis. Reflectivity data was normalized by the Fresnel Reflectivity (RF), which is the result of x-rays reflecting from an ideal interface. (R/RF) as a function of the wave vector transfer perpendicular to the interface (QZ) is fit to determine the electron density ρ(z) along the direction perpendicular to the surface. Two- and three-slab models were used to fit the reflectivity data to the model functional, as represented in
From the corresponding fitting of the XR data, we obtained values such as electron density, thickness, and interfacial roughness (Tables S4-S5). These parameters allowed us to compute surface concentration of peptide adsorbed to the air-water interface for each LBTs and LBTs/Tb3+ adsorbed layers to the air-aqueous interface from bulk solution. To compute the surface concentrations of molecules populating the air water interface we applied electron density and molar volume balances resulting in:
XFNTR is used to calculate the coverage of Tb3+ cations at the air-aqueous interface.
The integrated fluorescence intensity of Tb3+ Lα emission lines measured for values of QZ near the critical value (Qc) for the reference sample is represented in
The following Aspects are illustrative only and do not limit the scope of the present disclosure or the appended claims. It should be understood that the present technology can include any combination of any part or parts of any one or more of the following Aspects.
Aspect 1. A method, comprising: complexing a peptide-comprising surfactant (PEPS) to a rare earth element (REE) cation in a solution so as to form a PEPS-REE complex, the PEPS comprising a REE-binding region that preferentially binds to one or more REEs; and optionally recovering the REE cation.
The solution can include REE cations and non-REE cations. The solution can also include, for example, two or more types of REE cations.
Recovering can be accomplished by, effecting release of the cation from the binding loop; this can be accomplished with a stripping agent. Commonly used such agents can include, for example, include weak acid solutions (HCl) or structures that contain carboxylates like citrate; citric acid can also be used. Without being bound to any particular theory or embodiment, these can operate by either protonating the carboxylate groups (HCl) of the coordinating LBT sidechains or providing alternative carboxylate ligands to replace them (sodium citrate); citric acid can do both. It should be understood, of course, that the foregoing examples are not exhaustive.
One can make modifications to a PEPS's structure to alter charge on the PEPS while also preserving the PEPS's ability to bind selectively between lanthanide cations.
Shown in red (and labeled with A) are the coordinating ligands on this structure, green (label B) indicates non-coordinating negatively charged ligands, and blue (label C) is the positively charged N terminus. The free energy of binding of this peptide as measured by plate reader in our study is shown in the adjacent figure. La3+ is bound weakly, Tb3+ is bound most strongly, and there is a weak upturn in selectivity for lanthanides that are smaller and heavier than Tb3+.
Three variants of the PEPS sequence (
For example, for LBT1-4, we reduced the charge on the cation by amidation of the C terminus. On LBT1-3, we reduced the charge further by replacing the non-coordinating charged residue D11 with the polar but uncharged N. It should thus accordingly be understood that a user can replace or modify a non-coordinating residue.
While these changes reduce the affinity of the PEPS for the cation, they preserve the selectivity among them of the parent structure, as is apparent in the free energy of binding for the LBT1-3 peptide.
Aspect 2. The method of Aspect 1, wherein the REE-binding region comprises a lanthanide binding tag (LBT or LBT tag). An LBT can be a peptide sequence designed to mimic and enhance the lanthanide binding ability of the naturally occurring EF hand structural motif that is present in lanthanide-binding proteins (for example, lanmodulin) and calcium binding proteins (for example, calmodulin).
Atoms of an LBT that coordinate the cation in the parent peptide LBT1 are highlighted in
The exemplary table below shows example sequences characterized for terbium binding. This table shows that binding can be retained and/or enhanced through a wide variety of mutations, including fusion of the LBT terminus to the protein SUMO (Small Ubiquitin-like Modifier).
Binding between the PEPS and the REE can be characterized by, for example, coordination number. One can define the coordination number (CN) to be the number of oxygen atoms (both charged and uncharged) from the LBT peptide that directly interact with the cation (i.e., not counting coordinated waters). A structure that has a CN as low as 6 in the bulk that has adsorbed as LBT:REE complex with selectivity for one cation over another. Without being bound to any particular theory, a CN less than 6 may not in all instances have the synergistic responses within the complex to discriminate between cations. The PEPS can have a coordination number with the REE of, for example, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, or 15; the foregoing values are exemplary and are not limiting.
The CN and the spatial arrangement of the coordinated oxygens varies from cation to cation. The strength and number of these interactions work in a coordinated fashion to generate the affinity of the binding loop for the cation and the selectivity among the cations in bulk solution.
Without being bound to any particular theory, binding occurs in the bulk, followed by adsorption of LBT:REE complexes at the air/liquid interface. In some instances, the selectivity for a given REE can be higher in the bulk than at the interface, but this is not necessarily the case. The ratio of the selectivity for a given REE in the bulk to the ratio of the selectivity for that REE at the interface can be, for example, 20:1, 19:1, 18:1, 17:1, 16:1, 15:1, 14:1, 13:1, 12:1, 11:1, 10:1, 9:1, 8:1, 7:1, 6:1, 5:1, 4:1, 3:1, 2:1, 1.9:1, 1.8:1, 1.7:1, 1.6:1, 1.5:1, 1.4:1, 1.3:1, 1.2:1, or even 1:1.
In some instances, a PEPS:REE complex can be net-negative in charge. In some instances, a PEPS:REE complex can be net-neutral in charge. In some instances, a PEPS:REE complex can be net-positive in charge. The characteristics of the PEPS can be selected to achieve any of the foregoing charge characteristics.
Aspect 3. The method of Aspect 2, wherein the PEPS further comprises at least one additional hydrophobic moiety in addition to the LBT, the at least one additional hydrophobic moiety optionally comprising at least one of a non-polar peptide, an alkyl chain, and a hydrophobic hydrocarbon-containing structure.
The amino acids that define peptide sequences are commonly divided into four groups, i.e., non-polar, polar, positively charged and negatively charged.
The non-polar peptides that can impart hydrophobicity include glycine (G), alanine (A), valine (V), cysteine (C), proline (P), leucine (L), isoleucine (I), methionine (M), tryptophan (W), phenylalanine (F). Without being bound to any particular theory or embodiment, their addition can enhance the hydrophobic nature without disrupting the binding loop.
LLA (leucine-leucine-alanine) was used as an example terminal hydrophobic moiety because it is a repeat of the terminal hydrophobic domain in the LBTs in the literature. These termini were selected because they were shown to stabilize the loop binding structure; this is thought to be related to the ability of LLA to form an alpha helix structure. As discussed elsewhere herein, we added a repeated LLA to form the enhance the surface activity without disrupting the stabilization of the loop.
Other sequences of the non-polar peptides at the termini or the binding loop can be used. Such sequences can be guided by, e.g., rational selection, molecular simulation and/or machine learning. They can be selected to provide surface activity while preserving the ability of the binding loop to interact selectively with the cation in the bulk and to retain the cation at the interface.
As described elsewhere herein, there is another location in the binding region sequence that can contribute to the hydrophobic character or the bound peptide and cation. In one sequence provided herein, the W (tryptophan) in the linear sequence, located at position 7 in the current binding loop is the so-called “antenna” position for fluorescence studies. The antenna is excited and transfers its excitation to bound cation.
In some mutants, the mutants replace the hydrophobic tryptophan with the more hydrophobic non-natural amino acid acridone at this location; acridone also acts an antenna. Tryptophan and acridone are both hydrophobic; acridone is more hydrophobic that tryptophan. Without being bound to any particular theory or embodiment, this can promote adsorption of the PEPS. It should be understood, however, that it is not a requirement to have an “antenna” in the PEPS, and one can replace the tryptophan with a different amino acid that does not act as an antenna in the manner of tryptophan.
In addition to adjusting the amino acid selection and/or sequence, one can also append alkyl chains or other hydrophobic hydrocarbon containing structures at the termini outside of the binding loop to adjust the hydrophobicity of the PEPS. As a non-limiting example, the following peptide was shown to be surface active:
Other adjustments can also be made to modify the hydrophobicity of the PEPS. One binding loop can include the following polar amino acids: the negatively charged aspartic acid D, the polar asparagine N, negatively charged glutamic acid E. In this example, within the binding loop, the polar amino acids D1, N3, D5, E9, and E12 all participate in multidentate binding of the cation Tb3 in the loop. In addition, the backbone carbonyl group at the W7 participates in the binding.
Mutant PEPS can be designed using ML approaches and rational approaches in which hydrophilic peptides with either charged or polar moieties that participate in the multi-dentate binding of the loop domain and the cation are swapped to change selectivity in the bulk and to retain the cation at the interface. In performing such mutations at one or more locations on the peptide, one can exploit the polar uncharged amino acids serine (S), thrionine (T), tyrosine (Y), asparagine (N) and glutamine Qc and the charged amino acids, the positively charged amino acids lysine (K), arginine (R) and histidine (H) and the negatively charged amino acids aspartic acid (D) and glutamic acid (E).
One can also perform PEPS mutations of non-coordinating ligands to manage excess charge on binding. Without being bound to any particular theory or embodiment, one can create mutants to reduce the non-selective binding that one can hypothesize is caused by the charge of -2 on the peptide that is bound to a single Tb cation. This approach can include swapping of charged and uncharged polar peptides that do not directly participate in coordination of the cation, e.g., replacing the D11 with an N11.
One can also change the C-terminus to reduce negative charge on peptide. Peptides typically have an N terminus that is positively charged and a C terminus (a terminal carboxy group) that is negatively charged. The negatively charged C terminus is predicted in simulation to participate in network formation at the interface. To reduce this network formation, one can replace the terminal carboxy group, for example, with a terminal —CONH2. One can also replace the terminal carboxy group with other groups that reduce the negative charge of the peptide.
Further (and again without being bound to any particular theory or embodiment), the introduction of spacers between the binding loop and the hydrophobe can protect the binding loop from the environment of the interface. For example, repeated ethoxy groups (shown below) can be inserted between the binding loop and the terminal hydrophobic groups to increase the distance between the interface and the binding loop domain.
It should be understood, however, that ethoxy groups are illustrative only, as the spacer is not limited to ethoxy groups, and other spacers can be used instead or even with ethoxy groups.
One can modify LBT structures to promote adsorption at the interface. A trypotophan residue is hydrophobic, and can also confer surface activity to the native structure. Additional hydrophobes can further promote adsorption to the fluid interface or can be designed to promote self-assembly at the interface or in solution. The hydrophobes can be grafted onto the PEPS structure like alkyl chains or other hydrocarbon structures that can be appended to selected sites on the LBT. Alternatively, the hydrophobes can be amino acid residues with hydrophobic side chains strategically added to locations within the peptide sequence.
Hydrophobe addition may not be necessary if the hydrophobic:hydrophilic balance of the PEPS is too hydrophobic, for example, causing the PEPS lose their solubility in water. In that case, one can add amino acid residues that are charged; one can do so such that the overall complex structure is neutral or positive.
Aspect 4. The method of Aspect 3, wherein (a) at least one hydrophobic moiety is disposed at an end of the PEPS, (b) the PEPS comprises a spacer, the spacer optionally comprising an ethoxy group, disposed between the at least one hydrophobic moiety and the REE-binding region of the PEPS, or both (a) and (b).
Aspect 5. The method of any one of Aspects 1-4, wherein the PEPS selectively binds a first REE over a second REE. It should be understood, however, that the disclosed technology can be used to separate REEs from one another that are comparatively close in size. The disclosed technology can be used to separate a given REE from all other REEs that may be present in a sample.
Aspect 6. The method of Aspect 5, wherein the first REE is Tb and the second REE is La. Other exemplary REEs include Pr, Nd, Dy, and Ce.
Aspect 7. The method of any one of Aspects 1-6, further comprising effecting disposition of the PEPS-REE complex to an air/liquid interface. Such an interface can be an air/water interface or other air/aqueous interface.
Aspect 8. The method of Aspect 7, wherein the air/water interface is present at a bubble.
Aspect 9. The method of any one of Aspects 1-8, wherein the method is performed so as to give rise to an aggregate comprising a plurality of PEPS-REE complexes.
Aspect 10. The method of Aspect 9, wherein the aggregate comprises from 1 to 50 PEPS-REE complexes.
Aspect 11. The method of any one of Aspects 1-8, wherein the method is performed so as to give rise to a film comprising a plurality of PEPS-REE complexes at an air/liquid interface. Such an interface can be an air/water interface or other air/aqueous interface.
Aspect 12. The method of Aspect 11, wherein the film is thicker than a monolayer.
Aspect 13. The method of any one of Aspects 1-12, further comprising sparging bubbles through the solution under such conditions that the PEPS-REE cation complex adsorbs to a bubble; and recovering the REE cation.
Aspect 14. The method of any one of Aspects 1-13, wherein the PEPS is designed using at least one of artificial intelligence and a genetic algorithm.
Without being bound to any particular theory or embodiment, a non-limiting computational peptide design scheme can include the following aspects:
Aspect 15. The method of Aspect 14, wherein the artificial intelligence comprises machine learning.
Aspect 16. The method of any one of Aspects 14-15, wherein the PEPS is designed to preferentially bind one REE over at least one other REE.
Aspect 17. A composition, comprising: a complex comprising (i) a peptide-comprising surfactant (PEPS), the PEPS comprising at least one lanthanide binding tag (LBT) binding region, the at least one LBT comprising one or more residues arranged in a binding region that coordinates with an REE cation so as to form a PEPS-REE cation complex, and (ii) an REE cation, the REE cation of the complex coordinated with the binding region of the PEPS.
The composition can be a liquid; the composition can also be a bubble or even a foam. The PEPS-REE cation complex can be situated at an air/liquid interface. Thus a composition according to the present disclosure can be an air/liquid interface having disposed therealong at least one PEPS-REE cation complex. The interface can be an air/water interface; the interface can also be another air/aqueous interface.
As described elsewhere herein, the composition can be a bubble or even a foam. The composition can be, for example, an amount of liquid having disposed therein a plurality of bubbles, the bubbles having PEPS-REE complexes disposed along their boundaries, i.e., at the interface between air and liquid that defines the bubble.
Also as described elsewhere herein, the PEPS can include any one or more of (i) one or more hydrophobic moieties at an end of the PEPS, (ii) a spacer (which can comprise repeat units) disposed between the LBT and an end of the PEPS, e.g., between the LBT and one or more hydrophobic moieties disposed at an end of the PEPS, or both (i) and (ii).
Aspect 18. The composition of Aspect 17, wherein the composition comprises an aggregate comprising a plurality of PEPS-REE complexes.
Aspect 19. The composition of Aspect 17, wherein the composition comprises a film comprising a plurality of PEPS-REE complexes.
Aspect 20. The composition of any one of Aspects 17-19, wherein the PEPS is designed using at least one of artificial intelligence and a genetic algorithm.
Aspect 21. The composition of Aspect 20, wherein the artificial intelligence comprises machine learning.
Aspect 22. The composition of any one of Aspects 20-21, wherein the PEPS is designed to preferentially bind one REE over at least one other REE. As described herein, a PEPS can preferentially bind one REE over at least one other REE
A composition according to the present disclosure can include, for example, a first population of complexes that comprise PEPS1:REE1 and a second population of complexes that comprise PEPS2:REE2, wherein PEPS1 differs from PEPS2 and REE1 differs from REE2. The first population can be more numerous than the second population; alternatively, the second population can be more numerous than the first. The first and second populations can reside at a liquid/air interface, for example, at an air/water interface. Such an interface can be comprised in a bubble; the interface can also be comprised in a foam.
Similarly, the disclosed methods can include contacting a first PEPS and a second PEPS (wherein the first PEPS differs from the second PEPS) to a liquid that contains multiple REEs, one of which REEs can bind preferentially to the first PEPS as compared to the second PEPS, and another of which REEs can bind preferentially to the second PEPS as compared to the first PEPS). PEPS:REE complexes can then form, with the complexes then arriving at an air interface with the liquid. The REEs of the complexes can then, as described elsewhere herein, be recovered.
Aspect 23. A method, comprising application of at least one of molecular dynamics, artificial intelligence, and a genetic algorithm to design a PEPS that comprises an LBT and preferentially binds one REE over at least one other REE.
Aspect 24. The method of Aspect 23, further comprising training a supervising learning model to provide a quantitative predictive correlation between a molecular dynamics simulation features and an experimental binding affinity (AG).
Aspect 25. The method of any one of Aspects 23-24, further comprising application of a genetic algorithm and machine learning predictions to screen a plurality of candidate PEPS mutants that comprise an LBT.
As explained here, using LBTs to capture REEs provides a number of advantages. Some of these advantages include:
The present application claims priority to and the benefit of U.S. patent application No. 63/371,080, “Peptide Sequences At Air-Aqueous Interfaces For Lanthanide Recovery” (filed Aug. 11, 2022) and of U.S. patent application No. 63/400,695, “Peptide Sequences and Compositions At Air-Aqueous Interfaces For Lanthanide Recovery” (filed Aug. 24, 2022) All foregoing applications are incorporated herein by reference in their entireties for any and all purposes.
This invention was made with government support under DE-SC0022240 awarded by the Department of Energy. The government has certain rights in the invention.
Number | Date | Country | |
---|---|---|---|
63400695 | Aug 2022 | US | |
63371080 | Aug 2022 | US |