94
Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics Ph.D. Thesis V. Palermo University of Bologna March 2003 © All Rights Reserved

Creation of Nanometre-Scale Islands, Wires and Holes on ... · [2] From [3] Confinement Of Electrons In Quantum Corrals On A Metal Surface, Crommie MF, Lutz CP, Eigler DM, Science

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • Creation of Nanometre-Scale

    Islands, Wires and Holes on Silicon Surfaces

    for Microelectronics

    Ph.D. Thesis

    V. Palermo

    University of Bologna

    March 2003

    © All Rights Reserved

  • 2 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

  • V.Palermo 3

    Università degli studi di Bologna

    Creazione di isole, fili e fori nanoscopici su superfici di silicio per microelettronica

    Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    Tesi di Dottorato di Ricerca Scienze Chimiche

    XV ciclo

    Presentata da: Vincenzo Palermo

    Relatore:Prof. Alberto Ripamonti

    Co-Relatore: Dott. Derek Jones

    Coordinatore di Dottorato: Prof. Goffredo Rosini

    Marzo 2003

  • 4 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

  • V.Palermo 5

    OMNIA IN MENSURA, ET NUMERO ET PONDERE.

    Sapientiae Salomonis, 11:20

  • 6 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

  • V.Palermo 7

    1. Introduction: nanotechnology and the future of computers The motivations of nanotechnology research

    In 1965 an electronic engineer named Gordon Moore, one of the future founders of Intel,

    noted that the performance of computers and their complexity doubled every 18 months

    and foresaw that computer power would continue to grow exponentially over the following

    years.

    This prevision, quite provocative for its time, actually came about and gained the name of

    “Moore’s Law”, and continues to hold for the trends of today’s computer industry. Since

    1965, the number of transistors present in an integrated circuit (IC) has increased from

    several hundred to more than ten million, and the minimum size of transistor elements has

    shrunk from several millimetres to ≈130 nanometres (fig. 1.1).

    Devices of such tiny dimensions are actually fabricated using lithographic techniques,

    where light is passed through an optical mask to react with a photo-sensitive layer (resist)

    Fig.1.1. Evolution of the number of transistors present on commercial computers [1].

  • 8 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    on the silicon wafer. This resist is then selectively removed and used as a mask for

    processing the silicon surface (fig. 1.2, left). The maximum resolution attainable depends

    upon the wavelength used, and current technology is near to its intrinsic resolution limit.

    On the other hand, there is strong scientific and economic demand for further development

    in IC miniaturization, to obtain more powerful and complex computers. Besides every-day

    life applications, more powerful computers are fundamental for much scientific research,

    such as climate change tracking, genome sequencing, and fluid dynamics. Increased

    miniaturization is also fundamental for reducing power which has to be dissipated by the

    chips which run at progressively higher frequencies. Energy consumption by

    microelectronic devices is already an issue, and represents one of the main obstacles for

    the continuing growth in wireless communication (cell phones, portable computers, CD

    and DVD players, digital cameras, etc.).

    Thus, it is expected that new production methods, different from current lithographic ones,

    will be developed, methods which allow modification of a surface well below the 100 nm

    limit, and even down to single atom manipulation. Techniques such as Scanning

    Tunnelling Microscopy and Atomic Force Microscopy are already capable of moving

    single atoms (see fig. 1.2, right) but, unfortunately, building a working nanodevice in this

    way would take a very long time, and these techniques are difficult to apply to large scale

    production.

    Nowadays, thousands of researchers are working in the nanotechnology field towards a

    new generation of microelectronic devices. Several possible solutions are competing for

    tomorrow’s computer architecture, and there is still no clear winner. It is likely that the

    final solution will be the combined use of different techniques and components (including

    molecules, nanowires and nanodots) as they become available together with conventional

    Fig. 1.2. Left: conventional litographic process [2]. Right: atomic manipulation of iron atoms on copper [3].

  • V.Palermo 9

    silicon technology.

    Below is a brief summary of the most recent developments in nanotechnology and

    nanoscience.

    Actual trends in nanotechnology

    Perhaps, the most fascinating idea for nanodevice construction is to use one single

    molecule, working as a complete device. The first molecular diodes (i.e. molecules

    conducting current only in one direction) were created in 1997; in 1999, a molecular fuse

    and a molecular transistor were demonstrated, although there was no possibility of wiring

    these devices to external contacts. In April 2001, James Heath and his group at UCLA

    fabricated an array of overlapping crossbars, and placed a small molecule of rotaxane

    between each crossbar (fig. 1.3, left). This composite molecule is made up of two

    component parts, the main rod-like molecular axis and a mobile ring “threaded” on it like a

    bead on a necklace, and can function as a molecular switch. A working 16-bit memory

    circuit was constructed using these molecules. For a brief review of these works, see [4]. In

    June 2002, a single molecule transistor was built by connecting an organic molecule to two

    metal contacts; the molecule contained one or two atoms of a transition metal (cobalt or

    vanadium) forming the active region of the device, supported by an organic backbone [5].

    Fig.1.3. Working nanodevices. Left: schematic representation of rotaxane molecules between crossed nanowires [4]. Right: SEM image of semiconductor nanowires forming a small circuit [9].

  • 10 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    Another approach to nanodevice fabrication has become possible through the discovery of

    carbon nanotubes, which were observed for the first time in 1991 by a Japanese electron

    microscopist studying the material deposited during arc-evaporation synthesis of fullerenes

    [6]. They consist of a graphite-like carbon seamless cylinder, with a diameter of several

    nanometers and lengths of up to a millimetre. Carbon nanotubes are very stable, can

    behave as metals or semiconductors, and can host other molecular or ionic species thus

    modifying their electrical behaviour. In 2001, Avouris and his group reported the first

    circuit made with a single nanotube [7]. A few months later, Cees Dekker presented a

    nanotube-based transistor able to amplify an input signal by a factor of ten, and built

    several logic circuits using these nanotube transistors [8].

    One problem with carbon nanotubes is that it is very difficult to control their electronic

    properties, i.e., their metallic or semiconducting behaviour. An alternative to carbon

    nanotubes are semiconductor nanowires. Silicon nanowires can be made using a laser to

    vaporize the silicon together with a metal catalyst, like iron or gold. The vapour condenses

    in nanosized drops of silicon and metal, from which the wires slowly grow out as more

    silicon is adsorbed. In 2001, a group at Harvard University [9] created a transistor by

    crossing two different nanowires. After this, the same group arranged four nanowires in a

    noughts and crosses grid, creating something like a 4-bit memory (fig. 1.3, right). Even

    metallic nanowires, made of platinum and silver, can be used in a crossed configuration to

    store information [10].

    There are some issues common to all these new technologies, though. First, it is difficult to

    imagine these methods applied to large-scale production. Up to now, the insertion of a

    molecule between two electrodes is an occasional, lucky event, while nanotubes and wires

    have to be positioned on the surface, creating the appropriate contacts on them manually.

    The large-scale production of integrated circuits using these building blocks will not be

    straightforward and does not seem imminent.

    Another issue is of an economic and not a scientific or technological nature. Since 1965,

    the cost of IC manufacturing plants has sky-rocketed. If the cost of semiconductor

    production plants continues to rise exponentially, in a few years such plants will cost up to

    $20 billion. This is a sizeable investment, even for large companies such as IBM or Intel.

    For this reason, it is likely that IC companies will resist changing to completely new

    technologies, closing down their existing plants. As it is clear that silicon will remain the

  • V.Palermo 11

    fundamental raw material of the IC industry for the foreseeable future, nanotechnology

    developments for microelectronics will need to be silicon-compatible. In these early days

    of nanotechnology, the most valid approach would seem to be the addition of molecular

    functions to existing silicon technology – using the latter as a foundation on which to build.

    Fabrication of self-organised structures on silicon

    The possibility of using the phenomenon of atomic or molecular self-organization to create

    nanostructures on silicon has already been demonstrated. The clean silicon surface shows

    in some cases a high degree of order and complex surface reconstruction, as will be

    described later. Several different ordered structures form spontaneously on this surface,

    such as series of monatomic steps or boundaries between reconstructed areas. It has been

    demonstrated that it is possible to use these structures to fabricate ordered nanodots and

    nanolines on the surface [11]. More recently, well-defined nanometric patterns have been

    obtained with selective etching of silicon using nitric oxide [12].

    In this study the possibility of creating different types of nanostructures on the silicon

    surface is explored. Methods had to be developed which were:

    - Simple. They must not need complex masks or lithographic steps to create the

    structure, but rather exploit self-organisation phenomena;.

    - High resolution: the silicon surface modifications should be on a scalelength of below

    100 nm;

    - Fast: billions of nanostructures have to form over the whole surface simultaneously, to

    be compatible with large-scale production requirements;

    - Cheap: they must not require expensive equipment (such as e-beam lithography), but

    exploit simple chemical and/or physical treatments to produce nanostructures on the

    silicon surface.

    In Chapter 2, the main characteristics of silicon are described. Chapter 3 provides a

    summary of the techniques used for this research. Chapter 4 examines the chemical etching

    of silicon in different liquid environments, and the effects of this etching on the surface at a

    nanoscopic level, with the creation of nanoholes.

  • 12 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    Chapter 5 describes the growth in ultra-high vacuum (UHV) of nanoscopic voids and

    islands on the silicon surface, and the effect of surface oxide on this growth.

    Chapter 6 discusses the modification of silicon surfaces in UHV, following the adsorption

    of molecules and thermal heating, to produce nanoislands and nanolines on silicon.

    The overall conclusions of our work are summarised in Chapter 7.

    Finally we will give some conclusions based on the results obtained, and discuss possible

    applications of the methods developed.

    Bibliography

    [1] From www.intel.com

    [2] From www.sematech.org

    [3] Confinement Of Electrons In Quantum Corrals On A Metal Surface, Crommie MF,

    Lutz CP, Eigler DM, Science 262 (5131): 218-220 Oct 8 1993.

    [4] Molecules Get Wired, Service RF, Science 294 (5551): 2442-2443 Dec 21 2001.

    [5] Coulomb Blockade And The Kondo Effect In Single-Atom Transistors, Park J,

    Pasupathy AN, Goldsmith JI, Chang C, Yaish Y, Petta JR, Rinkoski M, Sethna JP,

    Abruna HD, Mceuen PL, Ralph DC, Nature 417 (6890): 722-725 Jun 13 2002 ; Kondo

    Resonance In A Single-Molecule Transistor, Liang WJ, Shores MP, Bockrath M, Long

    JR, Park H, Nature 417 (6890): 725-729 Jun 13 2002 ; Nanotechnology - Electronics

    And The Single Atom, De Franceschi S, Kouwenhoven L, Nature 417 (6890): 701-702

    Jun 13 2002.

    [6] Smallest Carbon Nanotube, Ajayan PM, Ijima S, Nature 358 (6381): 23-23 Jul 2 1992.

    [7] Carbon Nanotube Inter- And Intramolecular Logic Gates, Derycke V, Martel R,

    Appenzeller J, Nano Letters 1 (9): 453-456 Sep 2001.

    [8] Logic Circuits With Carbon Nanotube Transistors, Bachtold A, Hadley P, Nakanishi T,

    Dekker C, Science 294 (5545): 1317-1320 Nov 2001.

  • V.Palermo 13

    [9] Logic Gates And Computation From Assembled Nanowire Building Blocks, Huang Y,

    Duan XF, Cui Y, Lauhon LJ, Kim Kh, Lieber CM, Science 294 (5545): 1313-1317

    Nov 9 2001.

    [10] Formation And Disappearance Of A Nanoscale Silver Cluster Realized By Solid

    Electrochemical Reaction, Terabe K, Nakayama T, Hasegawa T, Aono M, Journal Of

    Applied Physics 91 (12): 10110-10114 Jun 15 2002.

    [11] Fabrication And Integration Of Nanostructures On Si Surfaces, Ogino T, Hibino H,

    Homma Y, Kobayashi Y, Prabhakaran K, Sumitomo K, Omi H, Accounts Of Chemical

    Research 32 (5): 447-454 May 1999.

    [12] Ultrafine And Well-Defined Patterns On Silicon Through Reaction Selectivity,

    Prabhakaran K, Hibino H, Ogino T, Advanced Materials 14 (19): 1418-1421 Oct 2

    2002.

  • 14 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

  • V.Palermo 15

    2. Silicon surfaces

    The name silicon (silicio in Italian) comes from the latin word silex. Amorphous silicon

    was first isolated by Berzelius, in 1824, by reaction of potassium with silicon tetrafluoride.

    Thirty years later, the first crystalline silicon was prepared. Silicon makes up 25% of

    earth’s crust, and is the second most abundant element after oxygen. Elemental silicon is

    not found in nature, occurring as silicon oxide (sand, quartz, amethyst, flint, etc.) or

    silicates (asbestos, clay, mica, etc.). Perhaps no other element and its compounds has such

    a wide range of uses. Silicon compounds such as sand and clay are used in the building

    industry, as refractory materials for high-temperature applications and for enamels and

    pottery. Silica is the main component of glass, silicon carbide is an important abrasive, and

    silicones are commonly used polymers and lubricants.

    Here, the most interesting use of silicon, of course, is for the production of

    microelectronics devices. For this application, silicon of high purity (99.9999%) and of

    high crystallinity is needed. Table 2.1 lists some of the physical characteristics of silicon.

    High purity polycrystalline silicon is produced by the reaction of gaseous trichlorosilane

    with hydrogen in a furnace. Then, to prepare a single-crystal of silicon, the so-called

    Czochralski method is commonly used.

    Polycrystalline silicon is melted in a quartz furnace at 1415°C in an argon atmosphere.

    Then, a seed of single-crystal silicon is lowered into contact with the melt, and slowly

    pulled out . In this way, the crystal grows and a crystalline cylindrical ingot several metres

    long is created from the initial seed.

    After cooling down, the ingot is sliced into thin silicon wafers. The wafer surfaces are

    polished using a counter-rotating lapping machine in an Al2O3 slurry until the surface is

    very flat and shiny, ready for the lithographic processes.

  • 16 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    Another way to obtain single crystal silicon is the Floating Zone (FZ) method, in which a

    silicon cylinder is slowly passed through a heating ring. The area inside the ring melts, and

    solidifies smoothly, crystallising as it comes out of the ring yielding a single silicon crystal.

    Microelectronic devices are built on the silicon surface, which is the surface of interest

    here. Unfortunately, silicon surfaces are normally quite dirty and uneven at the atomic

    scale. Atmospheric oxygen and humidity react with silicon surfaces creating a thin layer of

    oxide (called “native oxide”), which is usually irregular and full of defects. Different kinds

    of contaminants also adsorb onto the surface. These are usually small organic molecules

    and microscopic dust particles. A clean surface, on exposure to the atmosphere, is

    completely covered with gas molecules, in less than 10-9 seconds. If the pressure is

    reduced, let’s say to 10-6 mbar, this time increases to 1 second. This is the reason why to

    study a clean surface we have to work in UHV, at pressures below 10-10 mbar.

    The atoms in the silicon crystal have a diamond-like structure, each atom having 4 bonds

    in a tetrahedral sp3 arrangement, with bond angles of 109.47 degrees. At the crystal

    surface, some atoms will have non-bonding orbitals “dangling” in the vacuum, i.e. sp3

    orbitals with a lone electron which are highly reactive. These orbitals are known as

    dangling bonds. To minimize surface energy, the surface will reorganize, by decreasing

    the number of dangling bonds.

    Table 2.1 Physical data of silicon [1].

    Atomic Weight 28.09 Lattice constant (A) 5.43095 Crystal structure Face-centered cubic

    (diamond) Melting point 1415 °C

    Density (g/cm3) 2.328 Boiling point 2355°C Atoms/cm3 5.0E22 Minority carrier

    lifetime (s) 2.5E-3

    Dielectric Constant 11.9 Specific heat (J/g °C)

    0.7

    Breakdown field (V/cm)

    ~3E5 Thermal conductivity (W/cm °C)

    1.5

    Electron affinity, x(V)

    4.05 Vapour pressure (Pa) 1 at 1650°C, 1E-6 at 900° C

    Energy gap (eV) at 300K

    1.12 Reactivity Inert to acids. Attacked by halogens and alkaline

    solutions Intrinsic carrier

    conc. (cm-3) 1.45E10 Oxidation states +4, -4

    Intrinsic Debye Length (µm)

    24 Energy of a Si-Si bond (eV)

    2.32

    Intrinsic resistivity (Ω-cm)

    2.3E5

  • V.Palermo 17

    Dangling bond densities and positions, and thus the type of surface reconstruction, will

    depend upon crystal orientation as well as the temperature and kinetics of the system.

    Fig. 2.1 shows a drawing of the main faces of a silicon crystal. The angle α between any

    (11n) face and the (100) face can be calculated from: 2/cos 2 += nnα . The angle

    between any (11n) face and the (111) face can be calculated from:

    )2(3/)2(cos 2 ++= nnα .

    The chemistry and physics of the faces are very different; a brief description will be given

    for the most important orientations.

    Si (100)

    On the (100) surface, each atom has two Si-Si bonds connecting it to the bulk and two

    dangling bonds pointing outward. Surface energy is reduced by the dimerisation of the

    surface atoms through overlap interaction of one dangling bond per atom, forming rows of

    dimers aligned along the (110) direction. This is the well-known “2x1” reconstruction of

    this silicon surface. The symmetric dimers would make the silicon surface metallic, but to

    reduce surface stress the dimers tend to buckle, and the surface is thus semiconductive. It

    took several years to understand that the dimers are buckled, because at room temperature

    Fig. 2.1. Schematic view of the principal orientations of a silicon surface. Surface atoms are white, bulk atoms are black, dangling bonds are gray [2].

  • 18 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    they shift easily from one buckling direction to the other, thus appearing symmetric under

    STM observation. Fig. 2.2 shows an STM image of the 2x1 reconstructed surface.

    Even almost perfect (100) surfaces have a certain number of monoatomic steps and the

    dimer rows on atomic layers are aligned at 90° to those on adjacent layers. Dimer rows are

    thus perpendicular or parallel to the step. When the dimers on the upper side of the step

    are parallel to the step, the step is called SA; if they are perpendicular, the step is called SB

    Because of this symmetry, SA step edges will be smoother compared to the more broken,

    fragmented edges of the SB steps

    A common defect on the Si(100) surface is the presence of nickel contamination, which

    appear as missing dimers in STM images. This type of contamination is so critical that

    even if the silicon sample is only briefly brought into contact with stainless steel tools

    (tweezers, for example), the 2x1 reconstruction of the surface can be blocked.

    Silicon atoms can diffuse easily over the silicon surface as monomers and dimers

    especially at elevated temperatures. The anisotropy due to the 2x1 reconstruction causes a

    difference in the diffusion energies of adsorbates over the surface. Diffusion of these

    silicon species along dimer rows, for example, will be much easier. A list of diffusion

    energies for monomers and dimers is given below [2]:

    Diffusion on Si(100) 2x1 Ed (eV) Monomers along dimer rows 0.6 Monomers across dimer rows 0.85 Monomer formation energy 1.8 Dimer along dimer rows 1.1 Dimer across dimer rows 1.5 Dimer formation energy 2.6 Dimer binding energy 0.76 Vacancies along dimer rows 1.7 Vacancies across dimer rows 1.9

    SA

    SB

    Fig. 2.2. STM image of a 2x1reconstructed silicon surface,showing the dimer rows and steps,40x35 nm. Nickel-induced defectsare visible as dark spots. SA and SBsteps are indicated

  • V.Palermo 19

    So, the diffusion energy for both monomers and dimers is nearly 40% greater if they have

    to cross a dimer row. This difference reduces to ∼10% for vacancy diffusion.

    Si(113)

    The (113) surface can be imaged as a sequence of alternating (100) and (111)-like

    structures, with two and one dangling bonds on alternate atoms, respectively. Interest in the

    (113) surface is scientific, as it has been used to study the energetics of the (100) and (111)

    surfaces, as well as for surface adsorption experiments.

    Si(100) surfaces can easily develop (113) facets.

    Si(111)

    This surface, besides being the first one imaged with STM with atomic resolution, is one of

    the most studied, because it is the best cleavage face of silicon, and because it shows one

    of the most complex and elegant reconstructions in surface science.

    All Si-Si bonds in the silicon crystal are perpendicular to a (111) plane, so this face will

    have the lowest number of dangling bonds created per unit area. In fact, each Si atom on a

    (111) surface shows a single dangling bond oriented perpendicular to the surface, and

    bonded to three back atoms. These three bonds for each surface atom account for the great

    chemical and physical stability of the Si(111) surface. Surface energy is 0.09 eV Å-2,

    compared to 0.15 eV Å-2 for Si(100).

    For energy minimization, this surface reconstructs, forming a huge 7x7 lattice cell

    containing 102 atoms, described by the Dimer-Adatom-Stacking fault model (DAS). For a

    detailed description of cell structure, see fig. 2.3.

    The cell described by this model is very complex, being composed of three kinds of atoms:

    adatoms, rest atoms and corner hole atoms. Furthermore, a subsurface stacking fault is

    present in one half of the cell, making the two halves of the unit cell look different under

    STM (Fig. 2.4). It took 26 years of research to completely understand the exact structure of

    the 7x7 reconstruction.

    Cleaving a silicon crystal along a (111) plane produces a metastable 2x1 reconstruction;

    the 7x7 reconstruction is easily obtained by flashing at high temperature in UHV. At T >

  • 20 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    830°C, a disordered 1x1 phase covers the surface. Cooling down to 800°C leads to the

    formation of the 7x7 phase. If the cooling process is too rapid, small 7x7 domains nucleate,

    and a disordered 1x1 phase is preserved between domain boundaries.

    Si(110)

    Even though, as mentioned before, the (111) plane is the favoured cleaving plane of

    silicon, thin (100) commercial wafers will not break along this plane, because the angle

    between (100) and (111) is too far from 90° (see table 2.2). Instead, they will break along

    the (110) plane because it is perpendicular to the (100) surface. Each surface atom on

    Si(110) has a Si-Si bond pointing downward, one dangling bond pointing outward, and two

    Si-Si bonds parallel to the surface, in a zig-zag pattern (see fig. 2.1). Cleaved (110)

    surfaces are disordered but, upon annealing at high temperatures an ordered, complex 16x2

    Fig.2.3 Scheme of the 7x7 DAS model [2]. In each unit cell there are 9 dimers, 12 adatoms, and a stacking layer fault. The force driving this complex reconstruction is the minimization of dangling bonddensity. The DAS model shows the lowest number of dangling bonds (19) of all possiblereconstructions. 12 dangling bonds are at the adatoms, 6 at the rest atoms and 1 at the corner hole atom. This surface is metallic.

  • V.Palermo 21

    reconstruction takes place. The surface appears as a series of long ridges and valleys

    parallel to each other. Eventually, tilted facets of orientation (17 15 1) can form on this

    surface. The adsorption of Ge atoms on this surface leads to the formation of self-

    assembled nanowires [3].

    Table 2.2. Angles in degrees between different silicon faces [2].

    Orientation

    113

    110

    111

    100

    100

    25.24

    90.00

    54.74

    0

    111

    29.50

    35.26

    0

    110

    64.76

    0

    113

    0

    Fig. 2.4. STM image of a Si(111)surface, with 7x7 reconstruction. A unitcell with its adatoms is highlighted.Image size 13x13 nm.

  • 22 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    Bibliography

    [1] Weast RC, Handbook Of Chemistry And Physics (Chemical Rubber Co., Cleveland,

    1972)

    [2] Dabrowski J, Mussig H, Silicon Surfaces And Formation Of Interfaces (World

    Scientific Publishing, Singapore, 2000).

    [3] The Structure Of Clean And SiGe-Covered Si(110) Surfaces, Butz R, Luth H, Surface

    Science 365 (3): 807-816 Oct 1 1996

  • V.Palermo 23

    3. STM and other surface analysis

    techniques

    Scanning Tunneling Microscopy

    Since the invention of the optical microscope at the end of the 16th century, the possibility

    of examining surfaces at higher and higher magnification has fascinated mankind.

    Development of the technique continued and towards the end of the 19th century optical

    microscopes were as good as today's standard instruments. The physical limits of the

    wavelength of visible light (350-800nm) had been reached.

    In the 1920s de Broglie showed that electrons can behave like waves and the use of these

    particles for imaging with much higher resolution soon followed. Atomic resolution using

    this technique is only possible in the transmission mode with extremely carefully prepared

    samples.

    In 1982, using the peculiar properties of piezoelectric materials, Binnig and Rohrer brought

    a metallic tip very, very close to a silicon surface and scanned it across an extremly small

    area (fig.3.1). The tunneling of electrons from the tip into the sample or vice versa allowed

    them to obtain a local density of electronic states (LDOS) map of that surface. Although

    theory (which treated the extreme point of the tip as a sphere) then excluded the possibility

    of atomic resolution, following a tip "crash" into the surface, Binnig and Rohrer began to

    observe the LDOS with atomic resolution. For this discovery and their following work,

    they were awarded the Nobel Prize in Physics in 1986.

  • 24 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    The basic principle of STM is very simple. A metallic tip is scanned over a surface without

    making ohmic contact, and a tunneling current passes between the tip and the surface. An

    electronic circuit keeps this current constant by raising and lowering the tip during the

    scan. In this way, recording the tip height at each point a three-dimensional image of local

    density of electronic states (LDOS) of the surface can be obtained. To explain the

    extremely high resolution attainable by this simple technique, quantum theory is needed.

    According to classical physics, the current will flow between sample and tip only if they

    are in physical contact. If there is a vacuum gap between the two, the electrons will simply

    remain confined, for example, within the surface, without the possibility of passing into the

    tip.

    In quantum physics, however, the electrons have a certain probability of passing

    (tunneling) across the gap, appearing on the other side of the gap, in this way reaching the

    tip. It can be shown that the probability of an electron tunneling through a gap of thickness

    z is:

    kzep 22)0( −∝ ψ ; h

    φmk

    2= (1)

    where ψ(0) is the electron wavefunction at the surface-gap border, m is the electron

    mass=9.1x10-28 g, and φ is the work function of the metal (i.e. the energy required to

    remove an electron from that material. For silicon it is 4.8 eV). The tunneling current thus

    decays exponentially with z, and is extremely sensitive to topographical imperfections

    present on the scanned surface. A rough formula giving the current as a function of z is [1]:

    zFS eEVI

    φρ 025.1)( −∝

    Fig. 3.1. Binnig and Rohrer with the first STM. Image from IBM [2].

  • V.Palermo 25

    where ρs(EF) is the local density of states at the Fermi level on the given surface. For

    example, the formula predicts that, for silicon, an increase in tip-surface distance of 1 Å

    will give a 95% decrease in tunneling current.

    This huge dependence of tunneling current upon the distance allows detection even of the

    sub-nanometre changes in height given by the single atoms of which the surface is

    composed, and thus to resolve them in the LDOS images. Of course this description of the

    tunneling process is oversimplified and, for a more accurate one, the electronic states of

    the tip, of the sample and their interaction have to be taken into account. Fig. 3.2 shows a

    schematic representation of the interaction between tip and sample orbitals.

    The exponential decay of current with distance also yields high lateral resolution. If the tip

    is approximated as a sphere of radius R and the current passing at the minimum tip-sample

    distance is I0, then the current passing at a lateral distance x from this point will be:

    Rxk

    eII 22

    0

    2−

    =

    Assuming a tip radius of 100 nm, the current is concentrated in an area ∼1.5 nm wide at tip

    apex.

    Very sharp tips with even smaller curvature radii can be produced in several ways. Simple

    Fig.3.2. Schematic view of tip-sampleorbitals interaction. a) no interaction. b)equilibrium. c) sample positive. d) tippositive [1].

  • 26 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    tungsten wires mechanically cut are capable of obtaining atomic resolution on graphite in

    air, but for more disordered and rougher samples sharper and more reproducible tips are

    needed.

    STM tips are mostly made by electrochemically etching a W or Pt-Ir wire. The tips we

    used were prepared using methods based on the work of Fotino [3].

    A tungsten wire, 0.38 nm diameter, is immersed in a KOH solution(0.6M) with a thicker

    tungsten wire used as a counterelectrode. The cathodic and anodic reactions involved in the

    etching are:

    Cathode: 6H2O + 6 e- → 3H2(g) + 6 OH-

    Anode: W(s) + 8 OH- → WO42- + 4H2O + 6 e-

    A potential of 3V a.c. is applied to the tungsten, and the wire is immersed in the solution

    until a constant current of ∼100 mA is established. The etching rapidly removes metal,

    shaping the wire end as a sharp tip. When the potential reaches 11 V, the coarse tip etching

    is finished. The wire is then removed from the solution, carefully inserted into an

    insulating plastic tube leaving only the tip exposed, and re-immersed in the solution with

    the tip pointing upwards. A more gentle etching is thus made to reduce tip radius. Usually

    5 minutes etching at 0.7 V a.c. is used. In this configuration, very small hydrogen bubbles

    formed on the tip sides sliding upwards with a “honing” effect on the tip.

    This procedure yields extremely sharp and reproducible tips at the microscopic level.

    After the etching, the tip is thoroughly rinsed in ultrapure water, then dipped into

    concentrated HF to remove surface oxides and hydroxides [4]. The tip is dried with

    nitrogen, inserted into the UHV system and degassed overnight at ∼150°C.

    The possibility of measuring sub-nanometric distances would be useless without being able

    to control tip movement over such a minute scale. To scan the tip over the surface, a

    piezoelectric scanner is used. Piezos are usually made of an alloy of PbZrO3 and PbTiO3, a

    material which contracts or expands when a voltage is applied to it. The Omicron

    instrument used in our laboratory has three such piezo scanners for x,y, and z tip motion,

    allowing one to scan the tip over the surface with sub-Ångstrom precision (fig. 3.3).

    To isolate the instrument from ambient vibrations, the whole STM stage is suspended upon

    four springs, which eliminate all frequencies above 1 Hz, and surrounded by a crown of

  • V.Palermo 27

    copper wings and fixed magnets. Parasitic currents generated by the magnets into the

    copper wings contrast every movement of the stage, and efficiently block stage vibrations.

    The STM can be used not only to explore surface topography, but to measure the I/V

    characteristics of single atoms or molecules on the surface (Scanning Tunneling

    Spectroscopy or STS). Furthermore, it can be used to modify the surface with voltage

    bursts, digging into it or delicately moving single molecules or atoms over a surface [5]. It

    can be used in vacuum, in air and, with proper lateral isolation of the tip, even in liquids.

    A major drawback of STM is that it works only on conducting and semiconducting

    surfaces, and thus cannot be used on many surfaces of biological and chemical interest.

    Another instrument more suitable for these and other applications is the Atomic Force

    Microscope (AFM).

    Atomic Force Microscopy

    AFM was invented in 1986 by Binnig, Quate and Gerber after calculating the possibility of

    building a cantilever with a force constant of the same order of magnitude as that of a

    chemical bond.

    In AFM, a tip mounted on a microscopic cantilever (usually made of Si3N4, fabricated with

    optical lithography) is brought close to a surface. When the tip touches the surface, the

    cantilever is very slightly deflected upwards. The movement is measured by observing the

    Y-PIEZO

    SILICON SAMPLE

    Fig. 3.3. A picture of theSTM used for theexperiments. The tripodpiezo scanner is shown.

    Z-PIEZO

    X-PIEZO

    TIP

  • 28 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    deviation of a laser beam hitting the upper face of the cantilever. Fig. 3.4 provides a

    schematic view of the principle of AFM.

    The typical force constant of the cantilever varies from 0.0006 to 2 N/m, the typical

    resonance frequency is 3 to 120 kHz. The AFM tip can apply a force on the sample of up

    to 10-9 N. The AFM can be used on conductive or insulating surfaces, in vacuum, air or

    liquids. Furthermore, the tip can be modified to sense electrostatic potentials (electric force

    microscopy) or magnetic fields (magnetic force microscopy); it can even be functionalized

    with complex molecules such as proteins, to interact with biological surfaces.

    A drawback of the AFM is that the force it exerts can damage the surface under

    observation, especially if the sample is soft (as in the case of cell membranes, for

    example). This problem can be overcome using the instrument in tapping mode (where the

    tip does not move laterally during its brief contact with the surface) or in non-contact mode

    in which the tip oscillates above the surface during the scan and the changes in its

    frequency due to interaction with surface are monitored. The shifts in the oscillating

    frequency of the cantilever due to tip-sample interaction are then used for imaging the

    surface. In this mode, interaction of the tip with the surface is minimal, and soft samples

    can be imaged.

    STM and AFM are the main techniques used for this work. A brief description of other

    techniques used occasionally is given below.

    Fig. 3.4. Scheme of an atomic force microscope.

  • V.Palermo 29

    Low Energy Electron Diffraction (LEED)

    Electrons with energies in the 20-500 eV range are diffracted by a crystalline surface; the

    diffraction peaks are visualized on a fluorescent screen. This technique probes the long

    range order of the surface, up to a depth of several nanometres.

    X-Ray Photoelectron Spectroscopy (XPS)

    XPS allows both qualitative and quantitative chemical analysis of the elements present on

    or near the sample surface.

    An X-ray source is used to photoionize the atoms on a surface and produce photoelectrons.

    By measuring the kinetic energy of the photoelectrons, the binding energy of the electronic

    levels can be calculated. This energy will depend on the chemical environment of the

    surface atoms.

    Although the soft X-rays used penetrate to a depth of ~2000Å, the sampling depth of the

    technique is determined by the mean free path of the photoelectrons which allows their

    escape from only the first 10-100Å.

    Secondary Ion Mass Spectroscopy (SIMS)

    High and low energy ions (primary ions) are used to bombard a sample and remove surface

    atoms and ions. The ionic fragments removed (secondary ions) are then analysed by a mass

    spectrometer. The surface can be consumed during the measurement and profiles obtained,

    giving concentrations of the materials composing the sample at different depths (depth

    profiles).

    A popular variant of SIMS is TOF-SIMS. In this technique, the secondary ion masses are

    measured by a time-of-flight (TOF) measurement. The secondary ions generated by the

    bombarding primary ions are accelerated to a constant kinetic energy, and then move

    through a field-free space before they reach the detector, where their intensity is measured

    as a function of flight time. Since ions with different masses have different velocities at a

  • 30 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    given kinetic energy, the measured flight times of the ions can easily be converted to their

    masses. The static nature of this latter technique allows mass spectroscopy surface analysis

    with minimal damage to the surface.

    Bibliography

    [1] Chen CJ, Introduction To Scanning Tunneling Microscopy (Oxford University Press,

    Oxford, 1993)

    [2] From www.ibm.com

    [3] Tip Sharpening By Normal And Reverse Electrochemical Etching, Fotino M, Review

    Of Scientific Instruments 64 (1): 159-167 Jan 1993

    [4] A Convenient Method For Removing Surface Oxides From Tungsten STM Tips,

    Hockett LA, Creager SE, Review Of Scientific Instruments 64 (1): 263-264 Jan 1993

    [5] Confinement Of Electrons In Quantum Corrals On A Metal Surface, Crommie MF,

    Lutz CP, Eigler DM, Science 262 (5131): 218-220 Oct 8 1993

  • V.Palermo 31

    4. Surface modification of silicon in liquid.

    Nano-hole creation.

    Liquid treatments of silicon wafers are very common in the integrated circuit (IC)

    manufacturing industry. They are used to clean and improve surface uniformity, to create

    and etch protective oxide layers and to remove photo-resist layers.

    Crystalline silicon with its native oxide layer is very stable, and is resistant to many acids.

    It is easily attacked by hydrofluoric acid (HF) and alkaline solutions.

    The thin (~2 nm), passivating layer of native oxide (SiO2) is formed on exposure to the

    atmosphere. This surface layer contains many defects and contaminants, so it is usually

    chemically stripped and substituted with a better, chemically-formed protective oxide.

    The most common silicon cleaning procedure is the RCA method, named after the Radio

    Corporation of America [1]. It consists of two steps: in the first one, the surface is treated

    with a hot, alkaline solution (H2O:H2O2:NH4OH, 4:1:1), to remove particles from the

    surface; following this, a hot, acidic solution (H2O:H2O2:HCl, 4:1:1) is used to remove

    metal contamination. Other well-known cleaning methods are IMEC (a sequence of

    cleaning steps in H2O:O3 and dilute HF) or the Pirana etch (a hot 4:1 mixture of H2SO4 :

    H2O2).

  • 32 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    The standard RCA clean removes surface contaminants, etches the native oxide and

    oxidizes the silicon surface, leaving a uniform layer of silicon oxide which better protects

    the surface from further contamination.

    Etching with fluorine-based solutions

    Hydrofluoric acid is one of the most common reagents used in the treatment of silicon

    wafers, both in the research field and in industrial processes. A rapid dip in dilute HF is the

    simplest way to remove the native oxide from Si(100), and leaves the surface passivated by

    a layer of Si-H bonds. Because of the low polarization of Si-H bonds, the Si-H layer is

    stable even for several days, protecting the surface from contamination. It has often been

    assumed that this short etch does not significantly change the surface morphology of the

    silicon substrate[2], even though a prolonged dip in dilute HF leads to surface roughening

    [3].

    Although dilute HF roughens the Si(100) surface at the atomic scale [4,5], immersion in

    concentrated HF (49%) etches the surface oxide without attacking the Si surface,

    uncovering in this way the buried Si/SiO2 interface. The final, counter-intuitive result is

    that dilute HF etches the silicon while concentrated HF leaves the crystalline silicon

    untouched. [3]

    Etching Si with fluorine-containing solutions at different concentrations and pH can

    produce different morphologies, from rough surfaces to flat, nearly ideal Si-H terminated

    surfaces.

    Hessel et al. and Higashi et al. demonstrated in 1991 that very flat Si(111) surfaces can be

    obtained using 40% NH4F, while etching with HF always yields rough surfaces. The

    surface becomes smoother because the etchant rapidly attacks Si atoms at step borders,

    thus removing surface kinks and irregularities in a step-flow mechanism [6, 7]. Later on,

    even smoother and more perfect surfaces were obtained by removing oxygen from the

    solution, after it was discovered that oxygen dissolved in 40% NH4F initiates the formation

    of triangular etch pits. It was not possible to obtain flat surfaces by etching Si(100) with

    ammonium fluoride solutions, which leads to the formation of small 2x1 dimer-row

    reconstructed (100) terraces, together with (111) facets [8].

    This difference is caused by the different hydride terminations prevailing on the (100) and

    (111) faces. While the ideal Si(111)-H surface is monohydride terminated, the more

  • V.Palermo 33

    reactive dihydrides predominate on the Si(100)-H surface, making it more vulnerable to

    etching. The etching reaction is thus strongly anisotropic, etching (100) facets faster than

    (111), thus producing (111) microfaceting on Si(100) crystals.

    A more uniform Si(100) surface can, however, be prepared by etching at low pH with an

    HF/HCl mixture [9], or by using very dilute HF solutions and ultrapure water with low

    dissolved oxygen and carbon contents [10].

    Electrochemical etching can also be used, applying anodic or cathodic bias to the silicon,

    to obtain different morphologies [11]; by varying the potential, isotropic or anisotropic

    etching is observed. The aforementioned results show that, despite the simplicity of the

    reactants, fluoride etching of silicon is quite a complex reaction.

    Fig. 4.1. Chemical etching of silicon

    HO+H

    H HO

    H

    H

    H

    FSi

    Si

    SiSi

    +H2O

    -OH-+F-

    H2O H H

    Si H

    F OH

    H Si

    Si Si

    -H2

    +H2O

    H

    H

    H

    FSi

    Si

    SiSi

    H

    H

    H

    OHSi

    Si

    SiSi

    H

    H

    H

    H Si

    Si

    Si Si

    Etching mechanism of silicon

    HF rapidly dissolves the SiO2 passivating layer on silicon, leaving the surface almost

    completely hydrogenated [12]. After this, two different types of reactions etch the silicon

    simultaneously, one chemical and the other electrochemical [13]. The overall etching

    mechanism can be schematized in two stages (see Figure 4.1):

    i) Si-H bonds are substituted by Si-F or Si-OH bonds, creating a partial charge on the

    surface silicon atom and polarizing its Si-Si backbonds;

    ii) these polarised backbonds are then more easily attacked by HF or H2O. After

    rupture of the Si-Si bond, the atom is removed leaving behind new Si-H

    terminations, and the reaction can start again.

  • 34 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    These reactions take place, although at different rates, on both Si(100) and Si(111).

    Stage i) is usually the rate-determining step of the reaction, and the stability of the Si-H

    bonds depends upon the pH, the concentration of nucleophilic species in solution, and an

    eventual potential applied to the crystal.

    For pH >5, as in the case of concentrated NH4F solutions, reaction begins with attack by

    water to give Si-OH (step A→C). The -OH group is rapidly substituted by fluorine, with

    polarization of the underlying Si-Si bonds. These bonds are then easily attacked by water,

    the silicon atom being released into solution as HSiF(OH)2. The Si-OH → Si-F substitution

    is not fundamental for the reaction, and etching can proceed even for Si-OH terminated

    atoms, but XPS measurements showed the presence of a certain number of Si-F bonds

    remaining. Furthermore, fluorine seems to have a catalytic effect on Si-H substitution, as

    indicated by the dependence of the etch rate upon the F- concentration at least for pH

    values between 4 and 8.

    Si-F bonds can be easily removed by a water rinse. In the case of strongly alkaline

    solutions (pH=14), OH- groups act directly as nucleophiles, and no fluorine is needed to

    catalyze Si-H bond rupture.

    At pH

  • V.Palermo 35

    electrostatic interactions, by lowering the energy of the interaction step. After the

    formation of the Si-OH group, the reaction proceeds as shown in the scheme of fig. 4.1.

    Matsumura et al [4] proposed that not only water but HF2- molecules also play a major role

    in electrochemical etching of silicon, leaving on the surface Si-F terminated bonds which

    can be immediately attacked in an autocatalytic process (fig. 4.3).

    In the electrochemical reactions described above, an external potential is applied to the

    silicon crystal. The chemical and electrochemical reactions, in any case, take place

    simultaneously most of the time, with the chemical path predominating at high pH. Even

    when no external potential is applied to the silicon, partial electrochemical reactions can

    take place at different “cathodic” and “anodic” sites on the surface with an internal charge

    exchange which ensures neutrality [11]. This macroscopic silicon etching and hydrogen

    bubble formation can lead to visually observable patterns on the surface when Si(100) is

    immersed in ammonium fluoride, even without applying a potential.

    Si Si

    Si Si

    H

    H

    H

    F Si

    Si

    Si SiH

    F F-H+ -2e

    +HF2-

    H

    HF

    SiSi

    SiSi + F

    F F

    FH

    H

    H

    F -H+ -2e

    +HF2-

    Fig.. 4.3. Autocatalytic electrochemical etching of silicon by HF2- .

    Inhomogeneities on silicon surfaces caused by electrochemical reactions and charge

    transfer have been studied extensively, because they are of fundamental importance in the

    formation of porous silicon.

    Pore formation on silicon

    When Si(100) or Si(111) are etched under anodic bias in fluorine-based solutions,

    microscopic pores form on their surface. Several different morphologies of pores have

    been observed, with pore diameters ranging from 10 nm to several microns, with depths of

    several microns [14]. Pore shape is very variable too, ranging from ordered, straight pores

    to chaotic networks of branched pores (fig. 4.4). Porous silicon has been known since the

    fifties, but it was only in 1990 that interest in this material increased, following the

    discovery that porous silicon layers were able to emit bright, red light. This led to a large

    amount of research and now different classes of micropores can be reproducibly created,

  • 36 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    Fig. 4.4 Different types of Silicon micropores. From ref. [14].

    mostly for optical and micromachining applications. However, there is still no unified

    theory able to explain the nucleation and growth mechanism of all the different kinds of

    pores.

    We will give a short description of some of these theories; for more detail, see Parkhutik et

    al. [15].

    One model explains pore nucleation on the basis of physical processes such as hole

    positive charge migration, ion transport to the surface, and small perturbations on the

    silicon surface, modelled as Fourier components. The system is shown to be unstable, and

    some spatial frequencies that lead to pore nucleation evolve from the etching process.

    A second model focusses on stationary pore growth, without explaining the nucleation

    stage. According to this model, silicon dissolves preferentially at pore edges because h+

    charges are attracted by the stronger electric field present at these edges.

    A third class of models explains pore growth as a Diffusion Limited Aggregation (DLA)

    process, where the random walk of h+ charge carriers through the depleted layer present at

    the silicon-liquid interface controls pore shape.

    Finally, the model by Carstensen, Cristophersen and Foll [16], proposes that areas of the

    surface of some characteristic size LCO are etched by synchronized “current bursts” in the

    flow of h+ charges. These bursts dissolve silicon through cyclic stages of surface oxidation,

    oxide removal and hydrogen passivation. Areas where a burst has recently taken place are

    less passivated, and thus more likely to be etched again; in this way, the pore bottom

    continues to dissolve, while the pore walls are passivated and are thus less favourable areas

    towards current bursts.

  • V.Palermo 37

    EXPERIMENTAL RESULTS

    In the following sections, we will show some experimental results obtained from STM and

    AFM measurements of fluorine-treated Si(100) surfaces. In the first part, the results of

    mild etching, using concentrated and dilute HF solutions at low pH, are presented. In the

    second part, the results of etching at high pH using ammonium fluoride are presented, and

    the mechanism of pore formation discussed.

    Etching of Si(100) in dilute and concentrated HF

    Samples were cut from different areas of an 8-inch diameter p-doped silicon(100) wafer

    (10 Ω-cm) supplied by MEMC Electronic Materials. Each series of STM measurements

    was carried out over at least six different areas on at least two identical samples. Low

    Electron Energy Diffraction (LEED) was used to check the surface cleanliness of the

    samples before STM measurements.

    Table 4.1 summarizes the different treatments of each sample. After etching with

    electronic grade HF, each sample underwent a final rinse in Ultra-Pure Water (UPW,

    resistivity >18 MΩ-cm). Both the HF and the UPW were allowed to flow continuously

    over the sample surface. Some samples were not etched with HF at all but just washed with

    UPW to observe the morphology of the native oxide layer (~2 nm thick) covering the

    surface. All of these processes were carried out under nitrogen to limit reoxidation and the

    samples were then introduced from the nitrogen atmosphere directly into the vacuum

    chamber, and degassed overnight at ∼150°C before LEED and STM measurements.

    STM images were obtained from each sample using the same measurement parameters

    (sample bias 4 V, feedback current 1 nA, scan speed 800 nm s-1). These parameters

  • 38 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    provided a satisfactory level of reproducibility for all the samples. Measurements were

    made over an area of 500x500 nm (image size 500x500 pixel). Slope correction was

    carried out by subtracting row-wise and column-wise fitted slopes from the entire image,

    which gave better results than the simple subtraction of a fitted plane, especially for the

    rougher samples. Following slope correction, the rms roughness:

    ∑ −=xy

    hyxhN

    22 )),((

    and the 2-D Fourier transform:

    )(22

    2

    ),(4

    ),( vyuxixy

    eyxhvuF +∑∆= ππ

    were calculated for each image, where N2 is the number of pixels composing the image,

    h(x,y) is the surface height at each point, ∆ is the distance between points, h is the mean

    height and u, v are the spatial frequencies. The radial power spectrum PS(f) of the STM image data is obtained from the angular

    average of the squared Fourier transform with f 2 = u2 + v2.

    Fig. 4.5 shows the STM images obtained from the various samples. Sample A, still covered

    with its native oxide layer, shows an irregular surface, with RMS roughness of ~0.5 nm

    (see Table 4.1). Observing the sample with LEED gave no diffraction pattern, even at

    relatively high incident electron energies, because of the surface oxide coverage. After 1

    min etching in dilute HF (sample B), the morphology is similar to the original one, though

    Table 4.1. Sample treatments, average RMS roughness and integrated area of power spectra.

    Sample Treatment RMS roughness (nm)

    PS area (f < 0.1 nm-1)

    PS area (f > 0.1 nm-1)

    A Rapid dip in water 0.53 ± 0.13 3.66 0.32

    B 1 min. in HF 5% + 10 min. in water 0.51 ± 0.08 2.72 0.34

    C 30 min. in HF 5% + 10 min. in water 0.62 ± 0.08 10.64 0.38

    D 5 sec. in HF 49% + 10 min. in water 0.42 ± 0.04 1.83 0.22

  • V.Palermo 39

    Fig. 4.5 STM images of each group of samples, showing the topography of the silicon surface. A) noetching, original oxide surface. B) after 1 min etching in dilute HF. C) after 30 min etching in diluteHF. D) after dipping in concentrated HF. Grey scale indicates height of the surface, from lower (black) to higher (white). The images are 250x250 nm, i.e. representative portions of the images usedfor the roughness measurement and PSD analysis.

    some of the larger features have disappeared, and the image quality is better, maybe due to

    improved tunnelling due to the cleaner surface. The RMS roughness is comparable to that

    of the original surface. Clear diffraction patterns are visible using LEED, though at quite

    high energies (200 eV). After prolonged etching (sample C), the RMS roughness increases

    to 0.62 nm and a long-range corrugation is visible on the surface, even if the LEED pattern

    is good.

    The samples dipped in concentrated HF (D) reveal the bare Si/SiO2 interface, which has a

    disordered aspect and protrusions over a wide range of dimensions. The quality of the

    STM images of sample D is very good, probably due to the cleanliness of the surface,

  • 40 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    1 .E -0 2

    1 .E -0 1

    1 .E + 0 0

    1 .E + 0 1

    1 .E + 0 2

    1 .E + 0 3

    0 .0 0 0 .0 1 0 .1 0 1 .0 0f (1 /nm )

    nm^4

    A , a s re c e iv e d

    B , e tc h e d 1 min

    C, e tc h e d 3 0 min

    D, e tc h e d HF 4 9 %

    Fig. 4.6. Log-Log plot of the averaged power spectra of the STM images for all the samples.

    which gives a more stable tunnelling junction. The LEED pattern is excellent, showing

    clear diffraction peaks at energies as low as 37 eV, comparable to that obtained after high

    temperature cleaning in UHV.

    Fig. 4.6 shows the power spectra of the samples. The high frequency and low frequency

    areas of the power specturm are considered separately. Table 4.1 shows for each sample,

    together with the roughness, the integrated area of the power spectrum for the high and low

    frequency part.

    We first examine the differences between the samples in the low frequency part of the

    spectrum ( 0.15 nm-1. This effect could be due to the

  • V.Palermo 41

    improved surface cleanliness after etching with concentrated HF, which would give a more

    stable STM junction thus reducing the high frequency noise in the image.

    Fig. 4.5 and the analysis of the power spectrum of each sample shows that a rapid dip in

    HF removes the native oxide but does not lead to major changes in the morphology of the

    Si surface, its only effect being the removal of some of the larger features present on the

    original surface. Prolonged etching, on the other hand, increases the RMS roughness of the

    surface.

    Etching of Si(100) in concentrated ammonium fluoride and nano-hole creation

    Two different types of commercial p-doped Si(100) wafers (2Ω-cm and 10Ω-cm) from

    MEMC were used. Several different samples of 10x5 mm were immersed for 10 minutes

    in 40% electronic grade NH4F solution under agitation. Previous works used low

    temperatures or anodic potentials applied to the silicon to avoid gaseous hydrogen

    production and to obtain a uniform surface, but in our experiment we worked at room

    temperature to check the influence of hydrogen bubbles on surface morphology. During

    the etching, the stirring was sufficient to provide a uniform concentration of reagents over

    the whole sample surface, but not to mechanically remove the hydrogen bubbles from the

    silicon surface.

    After the etching, each sample was rinsed in ultra-pure water to remove any etching

    residues, and observed with STM, AFM and optical microscopy. The AFM measurements

    were made in air, while for STM measurements the samples were rapidly dried with

    nitrogen and inserted into the vacuum system to avoid surface reoxidation. After insertion

    into the vacuum, surface cleanliness was checked with LEED, and the surface morphology

    observed by STM. Parameters for STM measurements were sample bias +4 V, 1 nA

    current, 1.6 Hz scan rate. The images obtained were stable and reproducible over several

    days. Scan parameters for AFM were 20 nN force and 1 Hz scan rate.

    Some of the samples were cleaned with an RCA standard clean [1] before NH4F etching, to

    check the influence of possible surface contaminants on the final results. Eight different

    samples were prepared and more than sixty STM images of the samples were taken at

    different points of the various samples.

  • 42 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    (110)

    (110)

    Fig.4.7. a,b,c) STMtopographic images of differentetching morphologies. Eachimage is 500x500 nm. Z-ranges are 10, 10 and 18 nmrespectively. d) STM image of a bridgecreated by etching of the lowerlayers of silicon (black arrow).Image is 250x250x6 nm.

    After ~2 min of immersion in the solution, hydrogen bubbles become visible on the sample

    surface. The production is slow and the bubbles are quite stable on the sample without

    detaching. Thus, some areas of the surface are masked from the liquid etching action.

    STM observations (fig. 4.7) show that, at the nanometer scale, the surface is unevenly

    covered with holes of radii ranging from 10 to 200 nm, with depths of 2-4 nm. These holes

    have a wide range of different shapes and distributions. In most cases the surface was

    covered with a uniform distribution of round-shaped holes (fig.4.7a), indicating isotropic

    etching. The dimension and the density of the holes changed greatly from sample to

    sample and even over the surface of a single sample. In some cases the etching was

    anisotropic, yielding nearly square holes and layered structures, as shown in fig. 4.7b.

    Square holes have been previously observed in cases where the etching speed in the (110)

    direction is significantly smaller than in the (100) direction [17].

    Over these areas (fig 4.7b and especially 4.7c) it is clearly visible how, once the surface

    had been attacked, the reaction continued to preferentially remove atoms at step

    irregularities (kink atoms), straightening step edges. Eventually, the exposed underlying

    silicon was also attacked, and further holes created inside the previously etched larger

    ones. It was not possible to detect monatomic steps on this kind of surface. The smallest

    step height observed was ~1.5 nm, corresponding to several atomic layers. In the image

    shown in Figure 4.7a the etching was not strong and created only anisotropic holes on the

    surface. In fig. 4.7b and c the stronger etch proceeded laterally for several tens of

  • V.Palermo 43

    nanometres, leaving straight steps several tens of nanometers long, and roughly rectangular

    holes, as expected, given the structure of the (100) crystal face. In some cases, a

    significative underetch is observed, and the formation of suspended bridges and tunnels

    can be deduced from the STM images (fig. 4.7d)

    The formation of branched pores and suspended structures has been attributed, during pore

    formation, to diffusion limited aggregation effects, where the h+ charge carriers necessary

    for silicon etching have a higher probability of reacting at pore bottoms than reaching the

    upper part of the silicon surface. In the case of very deep pores, quantum wire effects have

    been invoked to explain the pore growth mechanism [15]. In our case, though, the pores

    formed were very shallow, the underetch depth being only a few nanometres on pores of

    ∼100 nm width. Thus, more than diffusion effects, the main contribution to the

    underetching process must come from anisotropic etching and some kind of autocatalytic

    reaction path, analogous to the one described by Matsumura et al. [4], with some areas of

    the silicon surface hydrogenated and thus less vulnerable to etching.

    Pre-treatment with RCA cleaning has no effect on the final morphology, and this seems to

    exclude pore nucleation being caused by presence of metallic or organic surface

    contaminants.

    The morphology and the distribution density of the pits was quite uniform over

    microscopic areas of the sample but changes were observed over the millimetre scale. This

    suggests that etching intensity is influenced by some large-scale parameter.

    Large-area measurements made with AFM or with an optical microscope (fig. 4.8),

    showed that the inhomogeneity of surface etching can be correlated with the masking

    action of the bubbles. While the fluoride dissolved the silicon, hydrogen bubbles formed

    by the reaction covered some areas of the surface thus blocking the etching over that area,

    generating macroscopic steps at the bubble-liquid border. As the reaction proceeded, more

    hydrogen accumulated and the bubble diameter increased, producing in this way a circular

    pattern of steps. The increase in bubble diameter was not continuous with time, otherwise a

    uniform surface slope gradient would have been obtained. The formation of this circular

    “etching staircase” indicates that the bubble growth was stepwise, the bubble accumulating

    more and more hydrogen without enlarging across the surface, until it relaxed, increasing

    its diameter stepwise and covering more silicon. The circular structures in fig. 4.8a are not

    co-axial and their asymmetry could derive from physical processes due to stirring or

    irregularities on the surface.

  • 44 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    a b Fig. 4.8. a) optical micrograph of etching patterns on Si(100) created by NH4F, 1.2x0.9 mm. b) AFM image of the circles border. xy range is 40x24 µm. z-range is 30 nm.

    The step structure was not destroyed by the etching even after the bubble detached from

    the surface but, on the contrary, the etching process seemed to be influenced by the

    presence of the step.

    Observing in detail a series of steps (fig. 4.8b), a quite deep trench is visible at the base of

    each step. A close-up image of a step and the corresponding line profile of the trench is

    shown in fig. 4.9. The trench is ∼5 nm deep with respect to the lower surface, compared to

    a step height of 2.2 nm.

    A similar structure has been recently obtained with electrochemical etching of p-type

    Si(100) in 4% HF, [18] in which a “current burst” etching model, previously described,

    was assumed for silicon dissolution. In that experiment, the trench was created at the

    border of silicon nitride masks, and began to grow after a nucleation stage. Preferential

    trench etching was along the (110) direction, and the trench growth was explained as an

    effect of mechanical stress induced by the nitride mask and of electric field enhanced

    dissolution which depended upon an external applied potential.

    While it is clear that in our system the gas bubbles have a masking effect similar to a

    classical, solid nitride mask, it is unlikely that hydrogen present on the surface can induce a

    significant stress in the silicon lattice as in the case of a nitride mask. Furthermore, no

    external field was applied to drive preferential etching at the trench site.

    It has been proposed [19] that the cathodic and anodic part of the etching reaction

    (hydrogen production and silicon oxidative etching, respectively) take place at different

    points on the surface, with a net charge transfer between the different areas. In this case,

    the highest reaction rates will correspond to the silicon area surrounding the bubble border,

    where a high number of positive charges will be available for the reaction. Furthermore, a

    sharp trench extending into the silicon crystal will be a preferential electrostatic attractor

  • V.Palermo 45

    Fig. 4.9. AFM image of the etched surface, showing a step created on the surface by bubblemasking. A stronger etching action is visible on the right side of the step, as well as theprotected area on the upper side of the step (indicated by the arrows). Image is 10x10 µm,z-range is 30 nm. The profile on the right is taken from the central area of the image.

    for the h+ charge carriers coming from other “cathodic areas” of the sample, either from

    other regions on the surface or from the back of the silicon chip [18].

    In the areas where the hydrogen bubble had detached and the surface was exposed to the

    etching, the reaction was not uniform in the neighbourhood of the steps. It is possible to

    observe (fig. 4.8b and 4.9) an area on the upper side of the step where less or even no

    etching at all seems to have taken place, as if the step was able to protect the surface from

    etching. While etching on the lower side with trench formation can be attributed to the

    presence of the bubble, the surface on the upper step side can be etched only after bubble

    detachment, so no masking effect can account for this result. However, a further

    preferential attraction of h+ charge carriers from the already formed trench can be

    hypothised, electrochemically shielding the surrounding area from further etching. If this is

    true, the shielding effect would be very strong, with a relatively shallow 5 nm-deep trench

    protecting an area of ∼1 µm parallel to the step.

    To summarise, the etching of Si(100) in NH4F is a complex process, in which different

    reaction paths, both chemical and electrochemical, co-exist. Hydrogen bubbles formed by

    the reaction act as a mask on the surface, and create etching paths and inhomogeneous

    etching of the surface. Different kinds of pores are observed on the surface, and in some

    cases the anisotropy of the process is so strong as to give square-shaped holes and

    underetching.

  • 46 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    The diffusion of h+ charge carriers in the crystal is one of the main rate-determining steps

    of the reaction, and leads to the formation of a deep trench immediately outside the bubble

    perimeter. These trenches act as charge collectors, and reduce the etching of the upper step

    surface in the proximity of the steps.

    Bibliography

    [1] The Evolution Of Silicon-Wafer Cleaning Technology, Kern W, Journal Of The

    Electrochemical Society 137 (6): 1887-1892 Jun 1990.

    [2] Spectroscopic Ellipsometry Studies Of HF Treated Si (100) Surfaces, Yao H, Woollam

    Ja, Alterovitz SA, Applied Physics Letters 62 (25): 3324-3326 Jun 21 1993; Influence

    Of HF-H2O2 Treatment On Si(100) And Si(111) Surfaces, Graf D, Bauermayer S,

    Schnegg A, Journal Of Applied Physics 74 (3): 1679-1683 Aug 1 1993; Kinetics Of

    Oxidation On Hydrogen-Terminated Si(100) And (111) Surfaces Stored In Air, Miura

    T, Niwano M, Shoji D, Miyamoto N, Journal Of Applied Physics 79 (8): 4373-4380

    Part 1 Apr 15 1996.

    [3] Hydrogen On Si - Ubiquitous Surface Termination After Wet-Chemical Processing,

    Pietsch GJ Applied Physics A-Materials Science & Processing 60 (4): 347-363 Apr

    1995; Structure Of The Stepped Si/SiO2 Interface After Thermal-Oxidation -

    Investigations With Scanning Tunneling Microscopy And Spot-Profile Analysis Of

    Low-Energy Electron-Diffraction, Pietsch GJ, Kohler U, Jusko O, Henzler M, Hahn

    PO, Applied Physics Letters 60 (11): 1321-1323 Mar 16 1992

    [4] Enhanced Etching Rate Of Silicon In Fluoride Containing Solutions At pH 6.4,

    Matsumura M, Fukidome H, Journal Of The Electrochemical Society 143 (8): 2683-

    2686 Aug 1996.

    [5] A Study Comparing Measurements Of Roughness Of Silicon And SiO2 Surfaces And

    Interfaces Using Scanning Probe Microscopy And Neutron Reflectivity, Crossley A,

    Sofield CJ, Goff JP, Lake ACI, Hutchings MT, Menelle A, Journal Of Non-Crystalline

    Solids 187: 221-226 Jul 1995.

  • V.Palermo 47

    [6] Step-Flow Mechanism Versus Pit Corrosion - Scanning-Tunneling Microscopy

    Observations On Wet Etching Of Si(111) By Hf Solutions, Hessel HE, Feltz A, Reiter

    M, Memmert U, Behm RJ, Chemical Physics Letters 186 (2-3): 275-280 Nov 8 1991

    [7] Comparison Of Si(111) Surfaces Prepared Using Aqueous-Solutions Of NH4F Versus

    HF, Higashi GS, Becker RS, Chabal YJ, Becker AJ, Applied Physics Letters 58 (15):

    1656-1658 Apr 15 1991

    [8] Wet Chemical Etching Of Si(100) Surfaces In Concentrated NH4F Solution -

    Formation Of (2x1)H Reconstructed Si(100) Terraces Versus (111) Faceting, Neuwald

    U, Hessel HE, Feltz A, Memmert U, Behm RJ, Surface Science 296 (1): L8-L14 Oct

    10 1993

    [9] Ideal Hydrogen Termination Of Si(001) Surface By Wet-Chemical Preparation, Morita

    Y, Tokumoto H, Applied Physics Letters 67 (18): 2654-2656 Oct 30 1995.

    [10] Atomic Structures Of Hydrogen-Terminated Si(001) Surfaces After Wet Cleaning

    By Scanning Tunneling Microscopy, Endo K, Arima K, Kataoka T, Oshikane Y, Inoue

    H, Mori Y, Applied Physics Letters 73 (13): 1853-1855 Sep 28 1998.

    [11] On The Potential-Dependent Etching Of Si(111) In Aqueous NH4F Solution

    Houbertz R, Memmert U, Behm RJ, Surface Science 396 (1-3): 198-211 Jan 20 1998

    [12] Etching Process Of SiO2 By HF Molecules, Hoshino T, Nishioka Y Journal Of

    Chemical Physics 111 (5): 2109-2114 Aug 1 1999.

    [13] Etching Mechanism And Atomic-Structure Of H-Si(111) Surfaces Prepared In

    NH4F, Allongue P, Kieling V, Gerischer H, Electrochimica Acta 40 (10): 1353-1360

    Jul 1995.

    [14] Pore Formation Mechanisms For The Si-HF System, Carstensen J, Christophersen

    M, Foll H, Materials Science And Engineering B-Solid State Materials For Advanced

    Technology 69: 23-28 Sp. Iss. Si Jan 19 2000

    [15] Porous Silicon - Mechanisms Of Growth And Applications, Parkhutik V, Solid-

    State Electronics 43 (6): 1121-1141 Jun 1999

    [16] Formation And Application Of Porous Silicon, Foll H, Christophersen M,

    Carstensen J, Hasse G, Materials Science & Engineering R-Reports 39 (4): 93-141 Nov

    1 2002

    [17] Pore Morphology And The Mechanism Of Pore Formation In N-Type Silicon,

    Searson PC, Macaulay JM, Ross FM, Journal Of Applied Physics 72 (1): 253-258 Jul 1

    1992.

  • 48 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    [18] Deep Electrochemical Trench Etching With Organic Hydrofluoric Electrolytes,

    Christophersen M, Merz P, Quenzer J, Carstensen J, Foll H, Sensors And Actuators A-

    Physical 88 (3): 241-246 Mar 5 2001.

    [19] Morphology Of Anodically Etched Si(111) Surfaces - A Structural Comparison Of

    NH4F Versus HF Etching, Houbertz R, Memmert U, Behm RJ, Journal Of Vacuum

    Science & Technology B 12 (6): 3145-3148 Nov-Dec 1994.

  • V.Palermo 49

    5. Surface modification of silicon in vacuum; void creation and oxide desorption

    The main reason for the huge success of silicon in the microelectronics industry is not due

    to its superior properties as a semiconductor. Other materials, for example germanium,

    have better qualities, such as higher mobility of charge carriers and lower noise levels,

    which would allow the construction of faster and higher performance devices.

    The widespread use of silicon, however, is mainly due to the outstanding characteristics of

    its oxide. Silicon dioxide (SiO2) is a very good electrical insulator, easy to form,

    chemically and thermally stable and is compatible with lithographic and metal deposition

    processes. Germanium oxide, on the contrary, is too reactive to be used.

    Even the use of Si(100) substrates for nearly all microelectronic devices is dictated by

    oxide quality. The (111) face of silicon crystal can be easily cleaved and flattened, and

    almost atomically perfect surfaces can be obtained with simple chemical procedures (as

    described above). But the density of interfacial defects is highest for Si(111)-SiO2

    interfaces, and lowest for Si(100)-SiO2 ones, so microchips will continue to be fabricated

    on Si(100) wafers.

    SiO2 (silica) is present in 95% of the earth’s minerals, in different allotropic forms such as

    quartz, tridymite, and cristobalite. In the bulk, each silicon atom is bonded to four oxygens,

    in a Si-O-Si tri-dimensional network. Si-O bonds are 0.16 nm long and form an angle

    ranging from 120° to 150°.

    Three typical intrinsic defects are present in SiO2. The so-called E’ centres are oxygen

    vacancies, with a hole localised on a silicon atom with only three Si-O bonds,

  • 50 Creation of Nanometre-Scale Islands, Wires and Holes on Silicon Surfaces for Microelectronics

    O3Si· +SiO3, Whereas the PR (peroxy radical) defects are holes trapped by a charged

    peroxy moiety, with a O3Si-O-O+ ·SiO3 structure. The NBOHC (non-bridging oxygen hole

    centres) derive from water or hydrogen contamination, and are schematized as O3Si- O- H-

    O-SiO3 .

    The atomic structure of the Si-SiO2 interface varies enormously. Local domains resembling

    the tridymite and the cristobalite structure of silica are present, but it seems that only 10%

    of the interface is ordered [1]. Far from the interface, the SiO2 bulk is completely

    disordered. The passage from bulk Si to stoichiometric SiO2 passes through a non-

    stoichiometric SiOx layer 0.7 nm thick.

    When a clean silicon surface is exposed to atmospheric oxygen, a thin, ∼2 nm thick layer

    of native oxide forms spontaneously which is usually removed and substituted with

    thicker, better quality oxide layers before further processing.

    Silicon is usually oxidised by thermal annealing, at temperatures between 800° and 1100°,

    in an atmosphere of pure O2 with some water eventually added to increase oxidation speed.

    Thermal oxides made in pure oxygen (dry oxides) grow more slowly than oxides produced

    in an oxygen-water atmosphere (wet oxides), but are usually of better quality.

    According to the Deal-Groove formula, the time t needed to grow an oxide of thickness X

    is given by [1]:

    1

    212−

    −−

    +=

    ABXBXt α

    where the constant B and B/A decrease exponentially with temperature as

    kTEexp ,

    with activation energies for dry oxidation of EB =1.23 and EB/A =2.0 eV respectively.

    EB is related to the diffusion of oxygen in silicon, while the value of EB/A is interpreted as

    the energy required to break a Si-Si bond. The exponent α is 1 for wet oxidation, and 0 for

    oxidation at high temperatures and low oxygen pressures. It has intermediate values for dry

    oxidation. This formula does not work well for low values of X, at the initial stages of

    oxidation, and usually empirical corrections are used.

    An interesting characteristic of silicon is that, at high temperature and in vacuum, oxygen

    can actually etch the silicon crystal giving gaseous products, and the oxidized layer present

    on the crystal becomes unstable (fig. 5.1) [2].

  • V.Palermo 51

    1.E-10

    1.E-09

    1.E-08

    1.E-07

    1.E-06

    1.E-05

    1.E-04

    1.E-03

    1.E-02

    1.E-01

    0.60.70.80.911.11.2

    1000/T (1/K)

    P (T

    orr)

    SiO2 + Si → 2SiO(g) Oxide decomposition

    Si+O2 → SiO2(s) oxide formation Fig.5.1. phase diagram of the oxygen-silicon system.

    Silicon oxidation, apart from the initial nucleation stages at the monolayer level, proceeds

    uniformly over the whole surface, with a planar reaction front moving from the surface

    into the bulk.

    If heated under low oxygen partial pressure (vacuum or inert atmosphere), SiO2 is known to

    decompose following the reaction:

    SiO2 + Si → 2SiO↑ (1)

    The reaction begins with nucleation at defect points on the Si/SiO2 interface and proceeds

    in a spatially inhomogeneous manner, with the formation of large voids on the oxide

    surface [3].

    Several studies have been made on the dynamics of void growth both on thick [4] and thin

    [5] layers of SiO2 . The process has been used to decorate otherwise unobservable defects at

    the Si/SiO2 interface [6], or to grow nanoislands of silicon on the void surface [7]. It has

    been suggested that the defects acting as nucleation centres could be metallic contaminants

    present on the native surface, which aggregate and catalyze SiO2 decomposition [8].

    After oxide desorption, the silicon surface is very rough. In particular, on Si(100) square

    islands are observed, several nanometres high, which act as ‘pinning sites’ on the motion of

    monatomic steps