Metal oxide adhesion

Last updated

The strength of metal oxide adhesion effectively determines the wetting of the metal-oxide interface. The strength of this adhesion is important, for instance, in production of light bulbs and fiber-matrix composites that depend on the optimization of wetting to create metal-ceramic interfaces. [1] The strength of adhesion also determines the extent of dispersion on catalytically active metal. [1] Metal oxide adhesion is important for applications such as complementary metal oxide semiconductor devices. These devices make possible the high packing densities of modern integrated circuits.

Contents

Oxide thermodynamics

Metal oxides are formed consistent with minimizing surface energy and minimizing system entropy. The formation reactions are chemical in nature, forming bonds between oxygen dimers and pure metals or metal alloys. The reactions are endothermic for transition metals and semi-metals. At isothermic and isobaric conditions at atmosphere, the probability for a free metal surface to bind an oxygen dimer via oxidation is a function of the partial pressure of oxygen, the surface energy between the crystal and the liquid or vapor phase (see heat of formation), and time.

In standard conditions, the determining factors for phase change are temperature and pressure. The idea here is that oxygen is making a phase change from gas to solid, and at the same time a bond is forming between oxygen and a metal. The instantaneous breaking of one bond and forming a different one required an energy contribution higher than the enthalpy of bond dissociation for molecular gaseous oxygen at 298K is +498.34 kJ/mol and is typically expressed as ∆Hf since it is also the heat of formation.

The majority of contributed entropy in the formation of metal-oxides is from O2(g). Gaseous oxygen molecules have high translation entropy, due to the excited vapor phase. This allows the transport of oxygen from the system to the interface or reaction surface. The change in entropy (ΔS) for oxidation is negative (exothermic) for semi-metals, transition metals, alkali earth metals and lanthanides/actinides. This fact is due to the elevated surface energy of an exposed pure metal and the ability of the tiny oxygen dimer to attract to high energy sites. The trend for oxide formation is that the reaction rate increases as atomic number increases.

Areas with elevated surface electron density will always oxidize preferentially, as is demonstrated beautifully in the formation of electro-anodized titanate. The formation of oxides is dominated by interactions between the Gibbs free energy surfaces of constituents. The intersections of Gibbs surfaces at a given temperature and pressure would be represented in 2D space as phase diagrams. In real world applications, Gibbs surfaces are subject to the additional dimension entropy. This third dimension constitutes a Cartesian coordinate space and the surface mapped out by the Gibbs energy for a given reaction gives a threshold energy needed for a phase transition. These values can be found in ASM library volumes, or online as the "standard heats of formation."

∆G=∆H-T∆S

standard state change of enthalpy is independent and thus the gradient of the change in Gibbs free energy as a function of temperature is linear. This dictates that an oxide becomes less thermodynamically stable with increasing temperature.

An important distinction between equilibrium wetting and non-equilibrium wetting is that the non-equilibrium condition occurs when a chemical reaction is taking place. This non-equilibrium wetting is an irreversible thermodynamic process that accounts for the changes of the chemical potential when forming a new boundary phase, such as an oxide.

Work of adhesion

The ideal work of separation Wsep is the reversible work needed to separate the interface into two free surfaces. [2] Important as a state function depending on the mechanical properties. [2] It is referred to as ideal because when the two free surfaces create an interface, the concentration of the interface will only be identical to the bulk at the instant the surface is created. In order to reach chemical equilibrium, the process of diffusion will take place which will increase any measurement of the work of separation. [2] The work of adhesion is the reversible free energy change for making free surfaces from interfaces. [2] It is represented by the equation:

where:

Wad is the work of adhesion

γm and γo are the respective surface energies of the metal and oxide

γmo is the surface energy between the two materials in contact

The following table gives some common metals and their corresponding surface energies. All the metals are face-centered cubic crystal structure and these surface energies correspond to the (100) surface plane.

MaterialSurface Energy
Al1.347
Pb0.377
Yb0.478
Cu2.166
Pd2.326
Ag1.200
Pt2.734
Au1.627

Oxide stability

Ellingham diagrams are generated according to the second law of thermodynamics and are a graphical representation of the change in the Gibbs free energy with respect to changing temperature for the formation of oxides.

Solid-gas interface

Structure

Real surfaces may be macroscopically homogeneous, but their microscopic heterogeneity plays a crucial role in the relationship between the metal and its oxide.

Transition metal oxides

A diagram showing different layers of transition metal oxides formed on the surface of a metal. Transition oxide.JPG
A diagram showing different layers of transition metal oxides formed on the surface of a metal.

Certain transition metals form multiple oxide layers that have different stoichiometric compositions. This is because the metal has multiple valency states with fewer or more electrons in the valence shell. These different valency states allow for multiple oxides to be formed from the same two elements. As the local composition of the material changes through diffusion of atoms, different oxides form as layers, one on top of another. The total adhesion in this situation involves the metal-oxide interface and oxide-oxide interfaces, which adds increasing complexity to the mechanics. [3]

Roughness

Increasing surface roughness increases the number of dangling bonds at the metal-oxide interface. The surface free energy of a crystal face is:


where:
E is the binding energy of the material

T is the temperature of the system

S is the surface entropy of the material

The binding energy favors a smoother surface that minimizes the number of dangling bonds, while the surface entropy term favors a rougher surface with increasing dangling bonds as the temperature is increased. [4]

Heterogeneity

A graph depicting the presence of thermodynamic defects, specifically lattice vacancies, as determined by the lowest potential free energy. As the temperature increases, the equilibrium concentration of lattice vacancies grows. Defect Energy.png
A graph depicting the presence of thermodynamic defects, specifically lattice vacancies, as determined by the lowest potential free energy. As the temperature increases, the equilibrium concentration of lattice vacancies grows.

Solid adsorption of an oxygen molecule depends on the heterogeneity of the substrate. Crystalline solid adsorption is dependent on the exposed crystal faces, grain orientations, and inherent defects because these factors provide adsorption sites with different steric configurations. Adsorption is largely determined by the reduction of Gibbs free energy associated with the exposed substrate.

Crystallographic orientation

A material's charge remains neutral when a surface is created by the law of charge conservation, but individual Bravais lattice planes, defined by their Miller indices, may be non-polar or polar based on their symmetry. A dipole moment increases the surface Gibbs free energy, but the greater polarizability of oxygen ions as compared to metals allows polarization to decrease the surface energy and thus increase the ability of metals to form oxides. Consequently, different exposed metal faces may adhere weakly to non-polar oxide faces, but be able to perfectly wet a polar face.

Defects

Surface defects are the localized fluctuations of surface electronic states and binding energies. Surface reactions, adsorption, and nucleation can be drastically affected by the presence of these defects. [5]

Vacancies
A figure demonstrating the surface adsorption of O2 based on surface reactions with the environment and the diffusion of the metal reactant to the reactive interface. Metal Oxide Reactions.png
A figure demonstrating the surface adsorption of O2 based on surface reactions with the environment and the diffusion of the metal reactant to the reactive interface.

Oxide growth is dependent upon the flux (diffusion) of either coupled or independent anions and cations through the oxide layer. [6] [7] Stoichiometric oxides have an integer ratio of atoms can only support coupled diffusion of anions and cations through the lattice migration of Schottky defects (paired anion/cation vacancies) or Frenkel defects (complete anion lattice with cation vacancies and interstitials). [6] [7] Non-stoichiometric oxide films support independent ion diffusion and are either n-type (extra electrons) or p-type (extra electron holes). Although there are only two valence states, there are three types: [6] [7]

  • cation excess (n-type)
  • anion deficit (n-type)
  • cation deficit (p-type)

Non-stoichiometric oxides most commonly have excess metal cations as a result of insufficient oxygen during the creation of the oxide layer. Excess metal atoms with a smaller radius than O2− anions are ionized within the crystal lattice as interstitial defects and their lost electrons remain free within the crystal, not taken by the oxygen atoms. The presence of mobile electrons within the crystal lattice significantly contributes to the conduction of electricity and the mobility of ions. [6]

Impurities

Impurity elements in the material can have a large effect on the adhesion of oxide films. When the impurity element increases the adherence of the oxide to the metal it is known as the reactive element effect or RE effect. Many mechanics theories exist on this topic. The majority of them attribute the increase in adhesion strength to the greater thermodynamic stability of the impurity element bonded with oxygen than the metal bonded to the oxygen. [2] [8] Inserting yttrium into nickel alloys to strengthen the oxide adhesion is an example of the reactive element effect.

Dislocations

Dislocations are thermodynamically unstable, kinetically trapped defects. Surface dislocations often create a screw dislocation when stress is applied. In certain cases, screw dislocations can negate the nucleation energy barrier for crystal growth. [5]

Oxide-support relationship

The adsorption of a monolayer of gas atoms is either commensurate or incommensurate. Commensurate adsorption is defined by having a crystal structure relationship between substrate-adsorbate layer that produces a coherent interface. Wood's notation is a description of the relationship between the simplest repeating unit area of the solid and adsorbate. The difference between the resulting commensurate interfaces can be described as an effect of misfit. The interfacial interaction can be modeled as the sg plus the stored elastic displacement energy due to lattice misfit. A large misfit corresponds to an incoherent interface where there is no coherency strain and the interface energy can be taken as simply the sg. In contrast, a small misfit corresponds with a coherent interface and coherency strain that results in the interfacial energy equivalent to the minimum sg. [9]

Strength of bond

A graph depicting the strength of forces as the stress required to debond materials increases. The positive trend seen shows as bonding increases, the force and stress required to debond the material does as well. Force diagram1.JPG
A graph depicting the strength of forces as the stress required to debond materials increases. The positive trend seen shows as bonding increases, the force and stress required to debond the material does as well.

The strength of the bond between the oxide and metal for the same nominal contact area can range from Pa to GPa stresses. The cause of this huge range stems from multiple phenomena dealing with at least four different types of adhesion. The main types of bonding that form adhesion are electrostatic, dispersive (van der Waals or London forces), chemical and diffusive bonding. As the adhesive forces increase, separation in crystalline materials can go from elastic debonding to elastic-plastic debonding. This is due to a larger number of bonds being formed or an increase in strength of the bonds between the two materials. Elastic-plastic debonding is when local stresses are high enough to move dislocations or make new ones. [10]

Solid-gas kinetics

When a gas molecule strikes a solid surface the molecule may either rebound or be adsorbed. The rate at which gas molecules strike the surface is a large factor in the overall kinetics of oxide growth. If there molecule is absorbed there are three potential outcomes. The surface interaction can be strong enough to dissociate the gas molecule into separate atoms or constituents. The molecule may also react with surface atoms to change its chemical properties. The third possibility is solid surface catalysis, a binary chemical reaction with a previously adsorbed molecule on the surface.

Dispersion

Most often it is kinetically favorable for the growth of a single oxide monolayer to be completed before the growth of subsequent layers. Dispersion in general can be modeled by:

where:

Ns is the number of atoms on the surface

Nt is the total number of atoms in the material

Dispersion is crucial to the growth of oxides because only atoms that are exposed to the interface can react to form oxides.

Diffusion

After the initial oxide monolayer is formed, new layers begin to build and the ions must be able to diffuse through the oxide in order to increase thickness of the oxide. The rate of oxidation is controlled by how fast these ions are able to diffuse through the material. As the thickness of the oxide increases, the rate of oxidation decreases because it requires the atoms to travel a further distance. This rate can quantified by calculating the rate of diffusion of vacancies or ions using Fick's first law of diffusion. [11]

where:
J is the flux and has units of mol·m2·s1
D is the diffusivity of the ions in the material

δC is the change in concentration of the material
δx is the thickness of the oxide layer

Solid surface catalysis

In 2007 the Nobel Prize in chemistry was awarded to Gerhard Ertl for the study of solid-gas interface molecular processes. One such process is the oscillatory kinetic catalysis. Oscillatory kinetic catalysis can be explained by different crystal surfaces favoring unmodified faces and reconstruction to reduce surface strain. The presence of CO can cause the reversal of surface reconstruction past a certain percent coverage. Once the reversal occurs, oxygen can be chemisorbed on the reverted surfaces. This produces an adsorption pattern with areas of surface coverage rich in CO and others O2. [12]

Driving force

The driving force of catalysis is determined by the difference between the unprimed equilibrium and the instantaneous interfacial free energies. [2]

See also

Related Research Articles

<span class="mw-page-title-main">Chemical reaction</span> Process that results in the interconversion of chemical species

A chemical reaction is a process that leads to the chemical transformation of one set of chemical substances to another. Classically, chemical reactions encompass changes that only involve the positions of electrons in the forming and breaking of chemical bonds between atoms, with no change to the nuclei, and can often be described by a chemical equation. Nuclear chemistry is a sub-discipline of chemistry that involves the chemical reactions of unstable and radioactive elements where both electronic and nuclear changes can occur.

<span class="mw-page-title-main">Crystal structure</span> Ordered arrangement of atoms, ions, or molecules in a crystalline material

In crystallography, crystal structure is a description of the ordered arrangement of atoms, ions, or molecules in a crystalline material. Ordered structures occur from the intrinsic nature of the constituent particles to form symmetric patterns that repeat along the principal directions of three-dimensional space in matter.

<span class="mw-page-title-main">Adsorption</span> Phenomenon of surface adhesion

Adsorption is the adhesion of atoms, ions or molecules from a gas, liquid or dissolved solid to a surface. This process creates a film of the adsorbate on the surface of the adsorbent. This process differs from absorption, in which a fluid is dissolved by or permeates a liquid or solid. While adsorption does often precede absorption, which involves the transfer of the absorbate into the volume of the absorbent material, alternatively, adsorption is distinctly a surface phenomenon, wherein the adsorbate does not penetrate through the material surface and into the bulk of the adsorbent. The term sorption encompasses both adsorption and absorption, and desorption is the reverse of sorption.

<span class="mw-page-title-main">Ionic compound</span> Chemical compound involving ionic bonding

In chemistry, an ionic compound is a chemical compound composed of ions held together by electrostatic forces termed ionic bonding. The compound is neutral overall, but consists of positively charged ions called cations and negatively charged ions called anions. These can be simple ions such as the sodium (Na+) and chloride (Cl) in sodium chloride, or polyatomic species such as the ammonium (NH+
4
) and carbonate (CO2−
3
) ions in ammonium carbonate. Individual ions within an ionic compound usually have multiple nearest neighbours, so are not considered to be part of molecules, but instead part of a continuous three-dimensional network. Ionic compounds usually form crystalline structures when solid.

<span class="mw-page-title-main">Surface energy</span> Excess energy at the surface of a material relative to its interior

In surface science, surface energy quantifies the disruption of intermolecular bonds that occurs when a surface is created. In solid-state physics, surfaces must be intrinsically less energetically favorable than the bulk of the material, otherwise there would be a driving force for surfaces to be created, removing the bulk of the material. The surface energy may therefore be defined as the excess energy at the surface of a material compared to the bulk, or it is the work required to build an area of a particular surface. Another way to view the surface energy is to relate it to the work required to cut a bulk sample, creating two surfaces. There is "excess energy" as a result of the now-incomplete, unrealized bonding between the two created surfaces.

<span class="mw-page-title-main">F-center</span> Type of crystallographic defect

An F center or Farbe center is a type of crystallographic defect in which an anionic vacancy in a crystal lattice is occupied by one or more unpaired electrons. Electrons in such a vacancy in a crystal lattice tend to absorb light in the visible spectrum such that a material that is usually transparent becomes colored. The greater the number of F centers, the more intense the color of the compound. F centers are a type of color center.

<span class="mw-page-title-main">Adhesion</span> Molecular property

Adhesion is the tendency of dissimilar particles or surfaces to cling to one another.

<span class="mw-page-title-main">Dangling bond</span>

In chemistry, a dangling bond is an unsatisfied valence on an immobilized atom. An atom with a dangling bond is also referred to as an immobilized free radical or an immobilized radical, a reference to its structural and chemical similarity to a free radical.

<span class="mw-page-title-main">Wetting</span> Ability of a liquid to maintain contact with a solid surface

Wetting is the ability of a liquid to maintain contact with a solid surface, resulting from intermolecular interactions when the two are brought together. This happens in presence of a gaseous phase or another liquid phase not miscible with the first one. The degree of wetting (wettability) is determined by a force balance between adhesive and cohesive forces.

<span class="mw-page-title-main">Contact angle</span> The angle between a liquid–vapor interface and a solid surface

The contact angle is the angle between a liquid surface and a solid surface where they meet. More specifically, it is the angle between the surface tangent on the liquid–vapor interface and the tangent on the solid–liquid interface at their intersection. It quantifies the wettability of a solid surface by a liquid via the Young equation.

Inner sphere complex is a type of surface complex that refers to the surface chemistry changing a water-surface interface to one without water molecules bridging a ligand to the metal ion. Formation of inner sphere complexes occurs when ions bind directly to the surface with no intervening water molecules. These types of surface complexes are restricted to ions that have a high affinity for surface sites and include specifically adsorbed ions that can bind to the surface through covalent bonding.

Kröger–Vink notation is a set of conventions that are used to describe electric charges and lattice positions of point defect species in crystals. It is primarily used for ionic crystals and is particularly useful for describing various defect reactions. It was proposed by F. A. Kröger and H. J. Vink.

In materials science, segregation is the enrichment of atoms, ions, or molecules at a microscopic region in a materials system. While the terms segregation and adsorption are essentially synonymous, in practice, segregation is often used to describe the partitioning of molecular constituents to defects from solid solutions, whereas adsorption is generally used to describe such partitioning from liquids and gases to surfaces. The molecular-level segregation discussed in this article is distinct from other types of materials phenomena that are often called segregation, such as particle segregation in granular materials, and phase separation or precipitation, wherein molecules are segregated in to macroscopic regions of different compositions. Segregation has many practical consequences, ranging from the formation of soap bubbles, to microstructural engineering in materials science, to the stabilization of colloidal suspensions.

<span class="mw-page-title-main">Ion</span> Particle, atom or molecule with a net electrical charge

An ion is an atom or molecule with a net electrical charge. The charge of an electron is considered to be negative by convention and this charge is equal and opposite to the charge of a proton, which is considered to be positive by convention. The net charge of an ion is not zero because its total number of electrons is unequal to its total number of protons.

Surface stress was first defined by Josiah Willard Gibbs (1839-1903) as the amount of the reversible work per unit area needed to elastically stretch a pre-existing surface. A suggestion is surface stress define as association with the amount of the reversible work per unit area needed to elastically stretch a pre-existing surface instead of up definition. A similar term called "surface free energy", which represents the excess free energy per unit area needed to create a new surface, is easily confused with "surface stress". Although surface stress and surface free energy of liquid–gas or liquid–liquid interface are the same, they are very different in solid–gas or solid–solid interface, which will be discussed in details later. Since both terms represent a force per unit length, they have been referred to as "surface tension", which contributes further to the confusion in the literature.

Adsorption is the accumulation and adhesion of molecules, atoms, ions, or larger particles to a surface, but without surface penetration occurring. The adsorption of larger biomolecules such as proteins is of high physiological relevance, and as such they adsorb with different mechanisms than their molecular or atomic analogs. Some of the major driving forces behind protein adsorption include: surface energy, intermolecular forces, hydrophobicity, and ionic or electrostatic interaction. By knowing how these factors affect protein adsorption, they can then be manipulated by machining, alloying, and other engineering techniques to select for the most optimal performance in biomedical or physiological applications.

Adsorption is the adhesion of ions or molecules onto the surface of another phase. Adsorption may occur via physisorption and chemisorption. Ions and molecules can adsorb to many types of surfaces including polymer surfaces. A polymer is a large molecule composed of repeating subunits bound together by covalent bonds. In dilute solution, polymers form globule structures. When a polymer adsorbs to a surface that it interacts favorably with, the globule is essentially squashed, and the polymer has a pancake structure.

Polymeric materials have widespread application due to their versatile characteristics, cost-effectiveness, and highly tailored production. The science of polymer synthesis allows for excellent control over the properties of a bulk polymer sample. However, surface interactions of polymer substrates are an essential area of study in biotechnology, nanotechnology, and in all forms of coating applications. In these cases, the surface characteristics of the polymer and material, and the resulting forces between them largely determine its utility and reliability. In biomedical applications for example, the bodily response to foreign material, and thus biocompatibility, is governed by surface interactions. In addition, surface science is integral part of the formulation, manufacturing, and application of coatings.

Transition metal oxides are compounds composed of oxygen atoms bound to transition metals. They are commonly utilized for their catalytic activity and semiconducting properties. Transition metal oxides are also frequently used as pigments in paints and plastics, most notably titanium dioxide. Transition metal oxides have a wide variety of surface structures which affect the surface energy of these compounds and influence their chemical properties. The relative acidity and basicity of the atoms present on the surface of metal oxides are also affected by the coordination of the metal cation and oxygen anion, which alter the catalytic properties of these compounds. For this reason, structural defects in transition metal oxides greatly influence their catalytic properties. The acidic and basic sites on the surface of metal oxides are commonly characterized via infrared spectroscopy, calorimetry among other techniques. Transition metal oxides can also undergo photo-assisted adsorption and desorption that alter their electrical conductivity. One of the more researched properties of these compounds is their response to electromagnetic radiation, which makes them useful catalysts for redox reactions, isotope exchange and specialized surfaces.

<span class="mw-page-title-main">Titanium biocompatibility</span>

Titanium was first introduced into surgeries in the 1950s after having been used in dentistry for a decade prior. It is now the metal of choice for prosthetics, internal fixation, inner body devices, and instrumentation. Titanium is used from head to toe in biomedical implants. One can find titanium in neurosurgery, bone conduction hearing aids, false eye implants, spinal fusion cages, pacemakers, toe implants, and shoulder/elbow/hip/knee replacements along with many more. The main reason why titanium is often used in the body is due to titanium's biocompatibility and, with surface modifications, bioactive surface. The surface characteristics that affect biocompatibility are surface texture, steric hindrance, binding sites, and hydrophobicity (wetting). These characteristics are optimized to create an ideal cellular response. Some medical implants, as well as parts of surgical instruments are coated with titanium nitride (TiN).

References

  1. 1 2 Peden, C; K.B. Kidd; N. D. Shinn (1991). "Metal/Metal-Oxide Interfaces: A surface science approach to the study of adhesion". Journal of Vacuum Science and Technology . 9 (3): 1518–1524. doi:10.1116/1.577656.
  2. 1 2 3 4 5 6 Finnis, M W (1996). "The theory of metal-ceramic interfaces". Journal of Physics: Condensed Matter. 8 (32): 5811–5836. doi:10.1088/0953-8984/8/32/003.
  3. Henrich, Victor; Cox P A (1996). The Surface Science of Metal Oxides. Cambridge University Press. ISBN   978-0-521-56687-2.
  4. Libbrecht, Kenneth (2005). "The physics of snow crystals" (PDF). Reports on Progress in Physics. 68 (4): 855–895. doi:10.1088/0034-4885/68/4/R03. S2CID   2648340.
  5. 1 2 Butt, Hans-Jurgen; Karlheinz Graf; Michael Kappl (2006). Physics and Chemistry of Interfaces. WILEY-VCH. pp. 167–169.
  6. 1 2 3 4 Kasap, S.O. (2006). Principle of Electron Materials and Devices. McGraw-Hill. pp. 73–75. ISBN   978-0-07-295791-4.
  7. 1 2 3 Behrens, Malte. "Solid State Kinetics" (PDF). Lecture Series. Fritz Haber Institute of the Max Planck Society Department of Inorganic Chemistry. Archived from the original (PDF) on 15 May 2011. Retrieved 1 June 2011.
  8. Pint, B A (2010). "Progress in Understanding the Reactive Element Effect Since the Whittle and Stringer Literature Review". Metals and Ceramics Division. 18 (18): 2159–2168.
  9. Johansson, Sven; Mikael Christensen; Goran Wahmstrom (2005). "Interface energy of semicoherent metal-ceramic interfaces". Physical Review Letters. 95 (22): 226108. doi:10.1103/PhysRevLett.95.226108. PMID   16384245.
  10. Gerberich, W W; M J Cordill (2006). "Physics of Adhesion". Reports on Progress in Physics. 69 (7): 2157–2204. doi:10.1088/0034-4885/69/7/R03.
  11. Rutter, N A. "Environmental Stability of Materials". Lecture Series. University of Cambridge. Archived from the original on 20 July 2011. Retrieved 6 June 2011.
  12. Ertl, Gerhard. "Chemical Processes on Solid Surfaces" (PDF). Nobel Prize. The Royal Swedish Academy of Sciences. Archived from the original (PDF) on 4 June 2011. Retrieved 6 June 2011.