Polymer brush

Last updated
Sample polymer brush Polymer brush (SCHEMATIC) V1.svg
Sample polymer brush

A polymer brush is the name given to a surface coating consisting of polymers tethered to a surface. [1] The brush may be either in a solvated state, where the tethered polymer layer consists of polymer and solvent, or in a melt state, where the tethered chains completely fill up the space available. These polymer layers can be tethered to flat substrates such as silicon wafers, or highly curved substrates such as nanoparticles. Also, polymers can be tethered in high density to another single polymer chain, although this arrangement is normally named a bottle brush. [2] Additionally, there is a separate class of polyelectrolyte brushes, when the polymer chains themselves carry an electrostatic charge.

Contents

The brushes are often characterized by the high density of grafted chains. The limited space then leads to a strong extension of the chains. Brushes can be used to stabilize colloids, reduce friction between surfaces, and to provide lubrication in artificial joints. [3]

Polymer brushes have been modeled with Molecular Dynamics, [2] Monte Carlo methods, [4] Brownian dynamics simulations, [5] and molecular theories. [6]

Structure

Polymer molecule within a brush. The drawing shows the chain elongation decreasing from the attachment point and vanishing at free end. The "blobs", schematized as circles, represent the (local) length scale at which the statistics of the chain change from a 3D random walk (at smaller length scales) to a 2D in-plane random walk and a 1D normal directed walk (at larger length scales). Tethered polymer chain.svg
Polymer molecule within a brush. The drawing shows the chain elongation decreasing from the attachment point and vanishing at free end. The "blobs", schematized as circles, represent the (local) length scale at which the statistics of the chain change from a 3D random walk (at smaller length scales) to a 2D in-plane random walk and a 1D normal directed walk (at larger length scales).

Polymer molecules within a brush are stretched away from the attachment surface as a result of the fact that they repel each other (steric repulsion or osmotic pressure). More precisely, [7] they are more elongated near the attachment point and unstretched at the free end, as depicted on the drawing.

More precisely, within the approximation derived by Milner, Witten, Cates, [7] the average density of all monomers in a given chain is always the same up to a prefactor:

where is the altitude of the end monomer and the number of monomers per chain.

The averaged density profile of the end monomers of all attached chains, convoluted with the above density profile for one chain, determines the density profile of the brush as a whole:

A dry brush has a uniform monomer density up to some altitude . One can show [8] that the corresponding end monomer density profile is given by:

where is the monomer size.

The above monomer density profile for one single chain minimizes the total elastic energy of the brush,

regardless of the end monomer density profile , as shown in. [9] [10]

From a dry brush to any brush

As a consequence, [10] the structure of any brush can be derived from the brush density profile . Indeed, the free end distribution is simply a convolution of the density profile with the free end distribution of a dry brush:

.

Correspondingly, the brush elastic free energy is given by:

.

This method has been used to derive wetting properties of polymer melts on polymer brushes of the same species [10] and to understand fine interpenetration asymmetries between copolymer lamellae [11] that may yield very unusual non-centrosymmetric lamellar structures. [12]

Applications

Polymer brushes can be used in Area-selective deposition. [13] Area-selective deposition is a promising technique for positional self-alignment of materials at a prepatterned surface.

See also

Related Research Articles

Navier–Stokes equations Equations describing the motion of viscous fluid substances

In physics, the Navier–Stokes equations are certain partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

Quantum decoherence Loss of quantum coherence

Quantum decoherence is the loss of quantum coherence. In quantum mechanics, particles such as electrons are described by a wave function, a mathematical representation of the quantum state of a system; a probabilistic interpretation of the wave function is used to explain various quantum effects. As long as there exists a definite phase relation between different states, the system is said to be coherent. A definite phase relationship is necessary to perform quantum computing on quantum information encoded in quantum states. Coherence is preserved under the laws of quantum physics.

Poissons equation Expression frequently encountered in mathematical physics, generalization of Laplaces equation.

Poisson's equation is an elliptic partial differential equation of broad utility in theoretical physics. For example, the solution to Poisson's equation is the potential field caused by a given electric charge or mass density distribution; with the potential field known, one can then calculate electrostatic or gravitational (force) field. It is a generalization of Laplace's equation, which is also frequently seen in physics. The equation is named after French mathematician and physicist Siméon Denis Poisson.

In the calculus of variations, a field of mathematical analysis, the functional derivative relates a change in a functional to a change in a function on which the functional depends.

Mass gap Energy difference between ground state and lightest excited state(s)

In quantum field theory, the mass gap is the difference in energy between the lowest energy state, the vacuum, and the next lowest energy state. The energy of the vacuum is zero by definition, and assuming that all energy states can be thought of as particles in plane-waves, the mass gap is the mass of the lightest particle.

In plasmas and electrolytes, the Debye length, is a measure of a charge carrier's net electrostatic effect in a solution and how far its electrostatic effect persists. With each Debye length the charges are increasingly electrically screened and the electric potential decreases in magnitude by 1/e. A Debye sphere is a volume whose radius is the Debye length. Debye length is an important parameter in plasma physics, electrolytes, and colloids. The corresponding Debye screening wave vector for particles of density , charge at a temperature is given by in Gaussian units. Expressions in MKS units will be given below. The analogous quantities at very low temperatures are known as the Thomas–Fermi length and the Thomas–Fermi wave vector. They are of interest in describing the behaviour of electrons in metals at room temperature.

Kerr–Newman metric

The Kerr–Newman metric is the most general asymptotically flat, stationary solution of the Einstein–Maxwell equations in general relativity that describes the spacetime geometry in the region surrounding an electrically charged, rotating mass. It generalizes the Kerr metric by taking into account the field energy of an electromagnetic field, in addition to describing rotation. It is one of a large number of various different electrovacuum solutions, that is, of solutions to the Einstein–Maxwell equations which account for the field energy of an electromagnetic field. Such solutions do not include any electric charges other than that associated with the gravitational field, and are thus termed vacuum solutions.

In Newton's theory of gravitation and in various relativistic classical theories of gravitation, such as general relativity, the tidal tensor represents

  1. tidal accelerations of a cloud of test particles,
  2. tidal stresses in a small object immersed in an ambient gravitational field.
Hofstadters butterfly Fractal describing the theorised behaviour of electrons in a magnetic field

In condensed matter physics, Hofstadter's butterfly is a graph of the spectral properties of non-interacting two-dimensional electrons in a perpendicular magnetic field in a lattice. The fractal, self-similar nature of the spectrum was discovered in the 1976 Ph.D. work of Douglas Hofstadter and is one of the early examples of modern scientific data visualization. The name reflects the fact that, as Hofstadter wrote, "the large gaps [in the graph] form a very striking pattern somewhat resembling a butterfly."

Local-density approximations (LDA) are a class of approximations to the exchange–correlation (XC) energy functional in density functional theory (DFT) that depend solely upon the value of the electronic density at each point in space. Many approaches can yield local approximations to the XC energy. However, overwhelmingly successful local approximations are those that have been derived from the homogeneous electron gas (HEG) model. In this regard, LDA is generally synonymous with functionals based on the HEG approximation, which are then applied to realistic systems.

In theoretical physics, scalar electrodynamics is a theory of a U(1) gauge field coupled to a charged spin 0 scalar field that takes the place of the Dirac fermions in "ordinary" quantum electrodynamics. The scalar field is charged, and with an appropriate potential, it has the capacity to break the gauge symmetry via the Abelian Higgs mechanism.

The method of image charges is a basic problem-solving tool in electrostatics. The name originates from the replacement of certain elements in the original layout with imaginary charges, which replicates the boundary conditions of the problem.

Cylindrical multipole moments are the coefficients in a series expansion of a potential that varies logarithmically with the distance to a source, i.e., as . Such potentials arise in the electric potential of long line charges, and the analogous sources for the magnetic potential and gravitational potential.

In 1927, a year after the publication of the Schrödinger equation, Hartree formulated what are now known as the Hartree equations for atoms, using the concept of self-consistency that Lindsay had introduced in his study of many electron systems in the context of Bohr theory. Hartree assumed that the nucleus together with the electrons formed a spherically symmetric field. The charge distribution of each electron was the solution of the Schrödinger equation for an electron in a potential , derived from the field. Self-consistency required that the final field, computed from the solutions, was self-consistent with the initial field, and he thus called his method the self-consistent field method.

Gravitational lensing formalism

In general relativity, a point mass deflects a light ray with impact parameter by an angle approximately equal to

In fluid dynamics, Luke's variational principle is a Lagrangian variational description of the motion of surface waves on a fluid with a free surface, under the action of gravity. This principle is named after J.C. Luke, who published it in 1967. This variational principle is for incompressible and inviscid potential flows, and is used to derive approximate wave models like the mild-slope equation, or using the averaged Lagrangian approach for wave propagation in inhomogeneous media.

A polymer field theory is a statistical field theory describing the statistical behavior of a neutral or charged polymer system. It can be derived by transforming the partition function from its standard many-dimensional integral representation over the particle degrees of freedom in a functional integral representation over an auxiliary field function, using either the Hubbard–Stratonovich transformation or the delta-functional transformation. Computer simulations based on polymer field theories have been shown to deliver useful results, for example to calculate the structures and properties of polymer solutions, polymer melts and thermoplastics.

The near-horizon metric (NHM) refers to the near-horizon limit of the global metric of a black hole. NHMs play an important role in studying the geometry and topology of black holes, but are only well defined for extremal black holes. NHMs are expressed in Gaussian null coordinates, and one important property is that the dependence on the coordinate is fixed in the near-horizon limit.

A depletion force is an effective attractive force that arises between large colloidal particles that are suspended in a dilute solution of depletants, which are smaller solutes that are preferentially excluded from the vicinity of the large particles. One of the earliest reports of depletion forces that lead to particle coagulation is that of Bondy, who observed the separation or "creaming" of rubber latex upon addition of polymer depletant molecules to solution. More generally, depletants can include polymers, micelles, osmolytes, ink, mud, or paint dispersed in a continuous phase.

Menter's Shear Stress Transport turbulence model, or SST, is a widely used and robust two-equation eddy-viscosity turbulence model used in Computational Fluid Dynamics. The model combines the k-omega turbulence model and K-epsilon turbulence model such that the k-omega is used in the inner region of the boundary layer and switches to the k-epsilon in the free shear flow.

References

  1. Milner, S. T. (1991). "Polymer Brushes". Science. 251 (4996): 905–14. Bibcode:1991Sci...251..905M. doi:10.1126/science.251.4996.905. PMID   17847384.
  2. 1 2 Chremos, A; Douglas, JF (2018). "A comparative study of thermodynamic, conformational, and structural properties of bottlebrush with star and ring polymer melts". J. Chem. Phys. 149 (4): 044904. Bibcode:2018JChPh.149d4904C. doi:10.1063/1.5034794. PMID   30068167.
  3. Halperin, A.; Tirrell, M.; Lodge, T. P. (1992). "Tethered chains in polymer microstructures". Macromolecules: Synthesis, Order and Advanced Properties. Advances in Polymer Science. Vol. 100/1. pp. 31–71. doi:10.1007/BFb0051635. ISBN   978-3-540-54490-6.
  4. Laradji, Mohamed; Guo, Hong; Zuckermann, Martin (1994). "Off-lattice Monte Carlo simulation of polymer brushes in good solvents". Physical Review E. 49 (4): 3199–3206. Bibcode:1994PhRvE..49.3199L. doi:10.1103/PhysRevE.49.3199. PMID   9961588.
  5. Kaznessis, Yiannis N.; Hill, Davide A.; Maginn, Edward J. (1998). "Molecular Dynamics Simulations of Polar Polymer Brushes". Macromolecules. 31 (9): 3116–3129. Bibcode:1998MaMol..31.3116K. CiteSeerX   10.1.1.465.5479 . doi:10.1021/ma9714934.
  6. Szleifer, I; Carignano, MA (1996). Tethered Polymer Layers. Adv. Chem. Phys. Vol. XCIV. p. 165. doi:10.1002/9780470141533.ch3. ISBN   978-0-471-19143-8.
  7. 1 2 Milner, S. T; Witten, T. A; Cates, M. E (1988). "A Parabolic Density Profile for Grafted Polymers". Europhysics Letters (EPL). 5 (5): 413–418. Bibcode:1988EL......5..413M. doi:10.1209/0295-5075/5/5/006.
  8. Milner, S. T; Witten, T. A; Cates, M. E (1989). "Effects of polydispersity in the end-grafted polymer brush". Macromolecules. 22 (2): 853–861. Bibcode:1989MaMol..22..853M. doi:10.1021/ma00192a057.
  9. Zhulina, E.B.; Borisov, O.V. (July 1991). "Structure and stabilizing properties of grafted polymer layers in a polymer medium". Journal of Colloid and Interface Science. 144 (2): 507–520. Bibcode:1991JCIS..144..507Z. doi:10.1016/0021-9797(91)90416-6.
  10. 1 2 3 Gay, C. (1997). "Wetting of a polymer brush by a chemically identical polymer melt". Macromolecules. 30 (19): 5939–5943. Bibcode:1997MaMol..30.5939G. doi:10.1021/ma970107f.
  11. Leibler, L; Gay, C; Erukhimovich, I (1999). "Conditions for the existence of non-centrosymmetric copolymer lamellar systems". Europhysics Letters (EPL). 46 (4): 549–554. Bibcode:1999EL.....46..549L. doi:10.1209/epl/i1999-00277-9.
  12. Goldacker, T; Abetz, V; Stadler, R; Erukhimovich, I; Leibler, L (1999). "Non-centrosymmetric superlattices in block copolymer blends". Nature. 398 (6723): 137. Bibcode:1999Natur.398..137G. doi:10.1038/18191.
  13. Lundy, Ross; Yadav, Pravind; Selkirk, Andrew; Mullen, Eleanor; Ghoshal, Tandra; Cummins, Cian; Morris, Michael A. (2019-09-17). "Optimizing Polymer Brush Coverage To Develop Highly Coherent Sub-5 nm Oxide Films by Ion Inclusion". Chemistry of Materials. 31 (22): 9338–9345. doi:10.1021/acs.chemmater.9b02856. ISSN   0897-4756.