Clove

Eugenol in cloves are used as a carminative ( prevent formation of gas in the gastrointestinal tract), to increase hydrochloric acid in the stomach and to improve peristalsis ( involuntary constriction and relaxation of the muscles of the intestine),  anthelmintic (used to destroy parasitic worms)

Cinnamon

Cinnamon have potent antiviral properties, improve glucose and lipids of people with type 2 diabetes due to cinnamic aldehyde

Chemical Composition of the human body

 

 

Composition  of the human  body

 

The composition of the human body can be looked at from several different points of view.

By mass, human cells consist of 65–90% water (H2O). Oxygen therefore contributes a majority of a human body’s mass. Almost 99% of the mass of the human body is made up of the six elements oxygen, carbon, hydrogen, nitrogen, calcium, and phosphorus.

 

About 0.85% is composed of only five elements: potassium, sulfur, sodium, chlorine, and magnesium. All are necessary to life. The remaining elements are trace elements, of which more than a dozen are thought to be necessary for life, or play an active role in health (e.g., fluorine, which hardens dental enamel but seems to have no other function).

Note that not all elements which are found in the human body in trace quantities play a role in life. Some of these elements are thought to be simple bystander contaminants without function (examples: caesium, titanium), while many others are thought to be active toxins, depending on amount (cadmium, mercury, radioactives). The possible utility and toxicity of a few elements at levels normally found in the body (aluminum) is debated. Trace amounts of cadmium and lead have had functions suggested, but are almost certainly toxic in amounts normally found in the body. There is evidence that one element normally thought a toxin (arsenic) is essential in ultratrace quantities, even in mammals. Some elements that are clearly used in lower organisms and plants (arsenic, silicon, boron, nickel, vanadium) are probably needed by mammals also, but in far smaller doses. Two halogens used abundantly by lower organisms (fluorine and bromine) are presently known to be used by mammals only opportunistically. However, a general rule is that elements found in active biochemical use in lower organisms are often eventually found to be used in some way by higher organisms.

 

 

Elemental  composition

The average 70 kg adult human body contains approximately 6.7 x 1027 atoms and is “composed of” 60 chemical elements. In this sense, “composed of” means that a trace of the element has been identified in the body. However, at the finest resolution, most objects on Earth (including the human body) contain measureable contaminating amounts of all of the 88 chemical elements which are detectable in nearly any soil on Earth. The number of elements thought to play an active positive role in life and augmentation of health in humans and other mammals, is about 24 or 25.[1]

 

The relative amounts of each element vary by individual. The numbers in the table are averages of different numbers reported by different references.

The human body is ~70% water, and water is ~11% hydrogen by mass but ~67% hydrogen by atomic percent.

 

 

Atomic number Element  

Percent  of

Mass[2][3][4][5]

 

Mass

(kg)[6]

Atomic percent Positive health role in mammals[7] Group
8 Oxygen 65 43 24 Yes (water, electron acceptor) /No (Reactive Oxygen

Species)

16
6 Carbon 18 16 12 Yes (organic compounds are hydrocarbon derivatives) 14
1 Hydrogen 10 7 63 Yes (e.g. water) 1
7 Nitrogen 3 1.8 0.58 Yes (e.g. DNA and amino acids) 15
20 Calcium 1.4 1.0 0.24 Yes (e.g. Calmodulin and Hydroxylapatite in bones) 2
15 Phosphorus 1.1 0.78 0.14 Yes (e.g. DNA and phosphorylation) 15
19 Potassium 0.25 0.14 0.033 Yes (e.g. Na+/K+-ATPase) 1
16 Sulfur 0.25 0.14 0.038 Yes (e.g. Cysteine and Methionine) 16
11 Sodium 0.15 0.10 0.037 Yes (e.g. Na+/K+-ATPase) 1
17 Chlorine 0.15 0.095 0.024 Yes (e.g. Cl-transporting ATPase) 17
12 Magnesium 0.05 0.019 0.0070 Yes (e.g. binding to ATP and other nucleotides) 2
26 Iron* 0.006 0.0042 0.00067 Yes (e.g. Hemoglobin) 8
9 Fluorine 0.0037 0.0026 0.0012 Yes/No (topically hardens teeth; toxic in higher amounts) 17
30 Zinc 0.0032 0.0023 0.00031 Yes (e.g. Zinc finger proteins) 12
14 Silicon 0.002 0.0010 0.0058 Yes (probable) 14
37 Rubidium 0.00046 0.00068 0.000033 No (?) 1
38 Strontium 0.00046 0.00032 0.000033 No (?) 2
35 Bromine 0.00029 0.00026 0.000030 No (?) 17
82 Lead 0.00017 0.00012 0.0000045 No (?) (toxic in higher amounts) 14
29 Copper 0.0001 0.000072 0.0000104 Yes (e.g. copper proteins) 11
13 Aluminium 0.000087 0.000060 0.000015 No(?) (toxic?) 13
48 Cadmium 0.000072 0.000050 0.0000045 No(?) (toxic in higher amounts) 12
58 Cerium 0.000057 0.000040   No  
56 Barium 0.000031 0.000022 0.0000012 No? (toxic) 2
50 Tin 0.000024 0.000020 6.0e-7 No(?) 14
53 Iodine 0.000016 0.000020 7.5e-7 Yes (e.g. thyroxine) 17
22 Titanium 0.000013 0.000020   No 4
5 Boron 0.000069 0.000018 0.0000030 Yes (probable) 13
34 Selenium 0.000019 0.000015 4.5e-8 Yes/No (toxic in higher amounts) 16
28 Nickel 0.000014 0.000015 0.0000015 Yes (e.g. urease) 10
24 Chromium 0.0000024 0.000014 8.9e-8 Yes (not confirmed) 6
25 Manganese 0.000017 0.000012 0.0000015 Yes (e.g. Mn-SOD) 7
33 Arsenic 0.000026 0.000007 8.9e-8 Yes (not confirmed). Toxic in higher amounts 15
3 Lithium 0.0000031 0.000007 0.0000015 Yes (not confirmed). Toxic in high amounts. Useful medically (mood stabilizer). 1
80 Mercury 0.000019 0.000006 8.9e-8 No (toxic) 12
55 Caesium 0.0000021 0.000006 1.0e-7 No 1
42 Molybdenum 0.000013 0.000005 4.5e-8 Yes (e.g. the molybdenum oxotransferases Xanthine oxidase and Sulfite oxidase 6
32 Germanium   0.000005   No (?) 14
27 Cobalt 0.0000021 0.000003 3.0e-7 Yes (e.g. vitamin B   )

12

9
51 Antimony 0.000011 0.000002   No (toxic) 15
47 Silver 0.000001 0.000002   No (toxic) 11
41 Niobium 0.00016 0.0000015   No 5
40 Zirconium 0.0006 0.000001 3.0e-7 No 4
57 Lanthanum 0.000137 8e-7   No  
52 Tellurium 0.000012 7e-7   No 16
31 Gallium   7e-7   No 13
39 Yttrium   6e-7   No 3
83 Bismuth   5e-7   No 15
81 Thallium   5e-7   No (toxic) 13
49 Indium   4e-7   No 13
79 Gold 0.000014 2e-7 3.0e-7 No 11
21 Scandium   2e-7   No 3
73 Tantalum   2e-7   No 5
23 Vanadium 0.000026 1.1e-7 1.2e-8 Yes (not confirmed) 5
90 Thorium   1e-7   No (toxic)  
92 Uranium 1.3e-7 1e-7 3.0e-9 No (toxic)  
62 Samarium   5.0e-8   No  
74 Tungsten   2.0e-8   No 6
4 Beryllium 5e-9 3.6e-8 4.5e-8 No (toxic) 2
88 Radium 1e-17 3e-14 1e-17% No (toxic) 2

  

*Iron = ~3 g in men, ~2.3 g in women

The elements needed for life are relatively common in the Earth’s crust, and conversely most of the common elements are necessary for life. An exception is aluminium, which is the third most common element in the Earth’s crust (after oxygen and silicon), but seems to serve no function in living cells. Rather, it is harmful in large amounts. Transferrins can bind aluminium.[8]

Periodic table highlighting dietary  elements[1]

 

H                                   He
Li Be                       B C N O F Ne
Na Mg                       Al Si P S Cl Ar
K Ca Sc   Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
Rb Sr Y   Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
Cs Ba La * Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
Fr Ra Ac ** Rf Db Sg Bh Hs Mt Ds Rg Cn Uut Fl Uup Lv Uus Uuo
     

*

 

Ce

 

Pr

 

Nd

 

Pm

 

Sm

 

Eu

 

Gd

 

Tb

 

Dy

 

Ho

 

Er

 

Tm

 

Yb

 

Lu

   
    ** Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr    


 

Composition  by molecule type

 

The composition can also be expressed in terms of chemicals, such as:

 

  • Water
  • Proteins – including those of hair, connective tissue, etc.
  • Fats (or lipids)
  • Apatite in bones
  • Carbohydrates such as glycogen and glucose
  • DNA
  • Dissolved inorganic ions such as sodium, potassium, chloride, bicarbonate, phosphate
  • Gases such as oxygen, carbon dioxide, nitrogen oxide, hydrogen, carbon monoxide, methanethiol. These may be dissolved or present in the gases in the lungs or intestines. Ethane and pentane are produced by oxygen free radicals.[9]
  • Many other small molecules, such as amino acids, fatty acids, nucleobases, nucleosides, nucleotides, vitamins, cofactors.
  • Free radicals such as superoxide, hydroxyl, and hydroperoxyl.

The composition of the human  body can be viewed on an atomic and molecular scale as shown in this article. The estimated gross molecular contents of a typical 20-micrometre human cell is as follows:[10]

 

Molecule Percent  of Mass Mol.Weight (daltons) Molecules Percent  of Molecules
Water 65* 18* 1.74e14* 98.73*
Other Inorganics 1.5 N/A 1.31e12 0.74
Lipids 12 N/A 8.4e11 0.475
Other Organics 0.4 N/A 7.7e10 0.044
Protein 20 N/A 1.9e10 0.011
RNA 1.0 N/A 5e7 3e-5
DNA 0.1 1e11 46* 3e-11

 

Water: Obviously the amount of water is highly dependent on the level of hydration. DNA: A human cell also contains mitochondrial DNA. Sperm cells contain less mitochondrial DNA than other cells. A mammalian red blood cell contains no nucleus and thus no DNA.

 

 

Materials and tissues

 

Body composition can also be expressed in terms of various types of material, such as: 

  • Muscle
  • Fat
  • Bone and teeth
  • Brain and nerves
  • Connective tissue
  • Blood – 7% of body weight.
  • Lymph
  • Contents of digestive tract, including intestinal gas
  • Urine
  • Air in lungs

 

Composition  by cell type

 

There are many species of bacteria and other microorganisms that live on or inside the healthy human body. In fact,

90% of the cells in (or on) a human body are microbes, by number[11][12] (much less by mass or volume). Some of these symbionts are necessary for our health. Those that neither help nor harm us are called commensal organisms.

  

References

 

[1]  Ultratrace minerals. Authors: Nielsen, Forrest H. USDA, ARS Source: Modern nutrition in health and disease / editors, Maurice E. Shils … et al.. Baltimore : Williams & Wilkins, c1999., p. 283-303. Issue Date: 1999 URI: (http://hdl.handle.net/10113/46493)

[2]  Thomas J. Glover, comp., Pocket Ref, 3rd ed. (Littleton: Sequoia, 2003), p. 324 (LCCN 2002-91021), which in turn cites Geigy Scientific

Tables, Ciba-Geigy Limited, Basel, Switzerland, 1984.

[3]  Chang, Raymond (2007). Chemistry, Ninth Edition. McGraw-Hill. pp. 52. ISBN 0-07-110595-6.

[4]  Distribution of elements in the human body (by weight) (http://www.daviddarling.info/encyclopedia/E/elbio.html) Retrieved on

2007-12-06

[5]  Zumdahl, Steven S. and Susan A. (2000). Chemistry, Fifth Edition. Houghton Mifflin Company. pp. 894. ISBN 0-395-98581-1.) [6]  J. Emsley, The Element, 3rd ed., Oxford: Clarendon Press, 1998.

[7]  Neilsen, cited (http://hdl.handle.net/10113/46493)

[8]  Mizutani, K.; Mikami, B.; Aibara, S.; Hirose, M. (2005). “Structure of aluminium-bound ovotransferrin at 2.15 Å resolution”. Acta

Crystallographica Section D Biological Crystallography 61 (12): 1636. doi:10.1107/S090744490503266X.

[9]  Douglas Fox, “The speed of life” (http://www.newscientist.com/article/mg18024195.500-the-speed-of-life.html), New Scientist, No

2419, 1 November 2003.

[10]  Freitas Jr., Robert A. (1999). Nanomedicine, (http://www.foresight.org/Nanomedicine/Ch03_1.html). Landes Bioscience. Tables 3–1 &

3–2. ISBN 1-57059-680-8. .

[11]  Glausiusz, Josie. “Your Body Is a Planet” (http://discovermagazine.com/2007/jun/your-body-is-a-planet). . Retrieved 2007-09-16. [12]  Wenner, Melinda. “Humans Carry More Bacterial Cells than Human Ones” (http://www.scientificamerican.com/article.


 

 

 

 

 

M.Tech U -V Fabrication

NANOCHEMISTRY

NANO FABRICATION-1

NANO FABRICATION-2,

Crystal Growth and Wafer Preparation

 

Unit-V   Fabrication:Crystal growth and wafer preparation, Defects, Clean room concept, Wafer cleaning techniques, Oxidation, Diffusion, Epitaxy, Ion implantation, Metallization, Lithography, Etching, Masking sequences and bipolar and MOS device fabrication process flow, Integration of unit process, Process modeling, Topological design rules, Passive device such as registers and capacitors and their non idealities, Fabrication of nanoelectronics structures.

 

M.Tech U-IV Applied chemistry of nanomaterials

Unit-IV   Applied chemistry of nanomaterials:Application to fundamenatal studies. Industrial applications: Photographic materials, Ceramic materials, Magnetic particles for recording media, Catalysts, Fuel cells electrocatalysis,Pigments, Nanostructured materials as new chemical reagents, Nanocomposite polymers, Fluids, inks and dyes, Block copolymers and dendrimers. Analytical and Environmental chemistry of nanoparticles.

 

Nanomaterials and their Applications

Nanomaterial Applications using Carbon Nanotubes

Applications being developed for carbon nanotubes include adding antibodies to nanotubes to form bacteria sensors, making a composite with nanotubes that bend when electric voltage is applied bend the wings of morphing aircraft, adding boron or gold to nanotubes to trap oil spills, include smaller transistors, coating nanotubes with silicon  to make anodes the can increase the capacity of Li-ion batteries by up to 10 times.

Nanomaterial Applications using Graphene

Applications being developed for graphene include using graphene sheets as electodes in ultracapacitors which will have as much storage capacity as batteries but will be able to recharge in minutes, attaching strands of DNA to graphene to form sensors for rapid disease diagnostics, replacing indium in flat screen TVs and making high strenght composite materials.

Nanomaterial Applications using Nanocomposites

Applications being developed for nanocomposites include a nanotube-polymer nanocomposite to form a scaffold which speeds up replacement of broken bones, making a graphene-epoxy nanocomposite with very high strenght-to-weight ratios, a nanocomposite made from cellulous and nanotubes used to make a flexible battery.

Nanomaterial Applications using Nanofibers

Applications being developed for nanofibers include stimulating the production of cartilage in damaged joints, piezoelectric nanofibers that can be woven into clothing to produce electricty for cell phones or other devices, carbon nanofibers that can improve the preformance flame retandant in funiture.

Nanomaterial Applications using Nanoparticles

Applications being developed for nanoparticles include deliver chemotherapy drugs directly to cancer tumors, resetting the immune system to prevent autoimmune diseases, delivering drugs to damaged regions of arteries to fight cardiovascular disease, create photocatalysts that produce hydrogen from water, reduce the cost of producing fuel cells and solar cells, clean up oil spills, water pollution and air pollution.

Nanomaterial Applications using Nanowires

Applications being developed for carbon nanotubes include using zinc oxide nanowires in a flexible solar cell, silver chloride nanowires to decompose organic molecules in polluted water, using nanowires made from iron and nickel to make dense computer memory – called “race track memo

 

Nanomaterials and their Applications

Nanomaterial Applications using Carbon Nanotubes

Applications being developed for carbon nanotubes include adding antibodies to nanotubes to form bacteria sensors, making a composite with nanotubes that bend when electric voltage is applied bend the wings of morphing aircraft, adding boron or gold to nanotubes to trap oil spills, include smaller transistors, coating nanotubes with silicon  to make anodes the can increase the capacity of Li-ion batteries by up to 10 times.

 

The properties of carbon nanotubes have caused researchers and companies to consider using them in several fields.  The following survey of carbon nanotube applications introduces many of these uses.

Carbon Nanotubes and Energy

Researchers at the University of Delaware have demonstarted increased energy density for capacitors whit the use of carbon nanotubes in 3-D structured electrodes.

Researchers at North Carolina State University have demonstrated the use of silicon coated carbon nanotubes in anodes for Li-ion batteries. They are predicting that the use of silicon can increase the capacity of Li-ion batteries by up to 10 times. However silicon expands during a batteries discharge cycle, which can damage silicon based anodes. By depositing silicon on nanotubes aligned parallel to each other the researchers hope to prevent damage to the anode when the silicon expands.

Researchers at Los Alamos National Laboratory have demonstrated a catalyst made from nitrogen-doped carbon-nanotubes, instead of platinum. The researchers believe this type of catalyst could be used in Lithium-air batteries, which can store up to 10 times as much energy as lithium-ion batteries.

Researchers at Rice University have developed electrodes made from carbon nanotubes grown on graphene with very high surface area and very low electrical resistance. The researchers first grow graphene on a metal substrate then grow carbon nanotubes on the graphene sheet. Because the base of each nanotube is bonded, atom to atom, to the graphene sheet the nanotube-graphene structure is essentially one molecule with a huge surface area.

Using carbon nanotubes in the cathode layer of a battery that can be produced on almost any surface. The battery can be formed by simply spraying layers of paint containing the components needed for each part of the battery.

Carbon nanotubes can perform as a catalyst in a fuel cell, avoiding the use of expensive platinum on which most catalysts are based. Researchers have found that incorporating nitrogen and iron atoms into the carbon lattice of nanotubes results in nanotubes with catalytic properties.

Carbon Nanotubes In Healthcare

Researchers are improving dental implants by adding nanotubes to the surface of the implant material. They have shown that bone adheres better to titanium dioxide nanotubes than to the surface of standard titanium implants. As well they have demonstrated to the ability to load the nanotubes with anti-inflammatory drugs that can be applied directly to the area around the implant.

Reseachers at MIT have developed a sensor using carbon nanotubes embedded in a gel; that can be injected under the skin to monitor the level of nitric oxide in the bloodstream. The level of nitric oxide is important because it indicates inflamation, allowing easy monitoring of imflammatory diseases. In tests with laboratory mice the sensor remained functional for over a year.

Researchers have demonstrated artificial muscles composed of yarn woven with carbon nanotubes and filled with wax. Tests have shown that the artificial muscles can lift weights that are 200 times heavier than natural muscles of the same size.

Nanotubes bound to an antibody that is produced by chickens have been shown to be useful in lab tests to destroy breast cancer tumors. The antibody-carrying nanotubes are attracted to proteins produced by one type of breast cancer cell. Once attached to these cells, the nanotubes absorb light from an infrared laser, incinerating the nanotubes and the attached tumor.

Researchers at the University of Connecticut have developed a sensor that uses nanotubes and gold nanoparticles to detect proteins that indicate the presence of oral cancer. Tests have shown this sensor to be accurate and it provides results in less than an hour.

Carbon Nanotubes and the Environment

Carbon nanotubes are being developed to clean up oil spills. Researchers have found that adding boron atoms during the growth of carbon nanotubes causes the nanotubes to grow into a sponge like material that can absorb many times it’s weight in oil. These nanotube sponges are made to be magnetic, which should make retrieval of them easier once they are filled with oil.

Carbon nanotubes can be used as the pores in membranes to run reverse osmosis desalination plants. Water molecules pass through the smoother walls of carbon nanotubes more easily than through other types of nanopores, which requires less power. Other researchers are using carbon nanotubes to develope small, inexpensive water purification devices needed in developing countries.

Sensors using carbon nanotube detection elements are capable of detecting a range of chemical vapors. These sensors work by reacting to the changes in the resistance of a carbon nanotube in the presence of a chemical vapor.

Researchers at the Technische Universität München have demonstrated a method of spraying carbon nanotubes onto flexible plastic surfaces to produce sensors. The researchers believe that this method could produce low cost sensors on surfaces such as the plastic film wrapping food, so that the sensor could detect spoiled food.

An inexpensive nanotube-based sensor can detect bacteria in drinking water. Antibodies sensitive to a particular bacteria are bound to the nanotubes, which are then deposited onto a paper strip. When the bacteria is present it attaches to the antibodies, changing the spacing between the nanotubes and the resistance of the paper strip containing the nanotubes.

Carbon nanotubes tipped with gold nanoparticles can be used to trap oil drops polluting water. Since the gold end is attracted to water while the carbon end is attracted to oil. Therefore the nanotubes form spheres surrounding oil droplets with the carbon end pointed in, toward the oil, and the gold end pointing out, toward the water.

Carbon Nanotubes Effecting Materials

Researchers  are developing materials, such as a carbon nanotube-based composite developed by NASA that bends when a voltage is applied. Applications include the application of an electrical voltage to change the shape (morph) of aircraft wings and other structures. This video from NASA gives you an idea of what a futuristic morphing aircraft might look like.

Researchers have found that carbon nanotubes can fill the voids that occur in conventional concrete. These voids allow water to penetrate concrete causing cracks, but including nanotubes in the mixstops the cracks from forming.

Researchers at MIT have developed a method  to add carbon nanotubes aligned perpendicular to the carbon fibers, called nanostiching. They believe that having the nanotubes perpendicular to the carbon fibers help hold the fibers together, rather than depending upon epoxy, and significanly improve the properties of the composite.

Avalon Aviation incorporated carbon  nanotubes in a carbon fiber composite engine cowling on an aerobatic aircraft to increase the strength to weight ratio. The engine cowling is highly stressed components in this aircraft, adding carbon nanotubes to the composite allowed them to reduce the weight without weakening the component.

Carbon Nanotubes and Electronics

Building transistors from carbon nanotubes enables minimum transistor dimensions of a few nanometers and the development of techniques to manufacture integrated circuits built with nanotube transistors.

Researchers at Stanford University have demonstrated a method to make functioning integrated circuits using carbon nanotubes. In order to make the circuit work they developed methods to remove metallic nanotubes, leaving only semiconducting nanotubes, as well as an algorithm to deal with misaligned nanotubes. The demonstration circuit they fabricated in the university labs contains 178 functioning transistors.

Other applications in this area include:

Carbon Nanotube Company Directory

Company Products
Nano Lab Functionalizied nanotubes and nanotube arrays
Bayer Material Science Carbon nanotubes
Cheap Tubes Carbon nanotubes

 

Nanomaterial Applications using Graphene

Applications being developed for graphene include using graphene sheets as electodes in ultracapacitors which will have as much storage capacity as batteries but will be able to recharge in minutes, attaching strands of DNA to graphene to form sensors for rapid disease

diagnostics, replacing indium in flat screen TVs and making high strenght composite materials.

The properties of graphene, carbon sheets that are only one atom thick, have caused researchers and companies to consider using this material in several fields. The following survey of research activity introduces you to many potential applications of graphene.

A Survey of Applications:

Hydrogen production without platimum. Researchers have demonstrated a catalyst made fromgraphene doped with cobalt can be used to produce hydrogen from water. The researchers at looking at this method as a low cost replacement for platimum based catalysts.

Lower cost of display screens in mobile devices. Researchers have found that graphene can replace indium-based electrodes in organic light emitting diodes (OLED). These diodes are used in electronic device display screens which require low power consumption. The use of graphene instead of indium not only reduces the cost but eliminates the use of metals in the OLED, which may make devices easier to recycle.

Lithium-ion batteries that recharge faster. These batteries use graphene on the surface of the anode surface. Defects in the graphene sheet (introduced using a heat treatment) provide pathways for the lithium ions to attach to the anode substate. Studies have shown that the time needed to recharge a battery using the graphene anode is much shorter than with conventional lithium-ion batteries.

Ultracapacitors with better performance than batteries. These ultracapacitiors store electrons on graphene sheets, taking advantage of the large surface of graphene to provide increase the electrical power that can be stored in the capacitor. Researchers are projecting that these ultracapacitors will have as much electrical storage capacity as lithium ion batteries but will be able to be recharged in minutes instead of hours.

Components with higher strength to weight ratios. Researchers have found that adding graphene to epoxy composites may result in stronger/stiffer components than epoxy composites using a similar weight of carbon nanotubes. Graphene appears to bond better to the polymers in the epoxy, allowing a more effective coupling of the graphene into the structure of the composite. This property could result in the manufacture of components with high strength to weight ratio for such uses as windmill blades or aircraft components.

Storing hydrogen for fuel cell powered cars. Researchers have prepared graphene layers to increase the binding energy of hydrogen to the graphene surface in a fuel tank, resulting in a higher amount of hydrogen storage and therefore a lighter weight fuel tank. This could help in the development of practical hydrogen fueled cars.

Lower cost fuel cells. Researchers at Ulsan National Institute of Science and Technology have demonstrated how to produce edge-halogenated graphene nanoplatelets that have good catalytic properties. The researchers prepared the nanoplatelets by ball-milling graphene flakes in the presence of chlorine, bromine or iodine. They believe these halogenated nanoplatelets could be used as a replacement for expensive platinum catalystic material in fuel cells.

Low cost water desalination: Researchers have determined that graphene with holes the size of a nanometer or less can be used to remove ions from water. They believe this can be used to desalinate sea water at a lower cost than the reverse osmosis techniques currently in use.

Lightweight natural gas tanks: Researchers at Rice University have developed a composite material using plastic and graphene nanoribbons that block the passage of gas molecules. This material may be used in applications ranging from soft drink bottles to lightweight natural gas tanks.

More efficient dye sensitized solar cells. Researchers at Michigan Technological University have developed a honeycomb like structure of graphene in which the graphene sheets are held apart by lithium carbonate. They have used this “3D graphene” to replace the platinum in a dye sensitized solar cell and achieved 7.8 percent conversion of sunlight to electricity.

Electrodes with very high surface area and very low electrical resistance. Researchers at Rice University have developed electrodes made from carbon nanotubes grown on graphene. The researchers first grow graphene on a metal substrate then grow carbon nanotubes on the graphene sheet. Because the base of each nanotube is bonded, atom to atom, to the graphene sheet the nanotube-graphene structure is essentially one molecule with a huge surface area.

Lower cost solar cells: Researchers have built a solar cell that uses graphene as a electrode while using buckyballs and carbon nanotubes to absorb light and generate electrons; making a solar cell composed only of carbon. The intention is to eliminate the need for higher cost materials, and complicated manufacturing techniques needed for conventional solar cells.

Transistors that operate at higher frequency. The ability to build high frequency transistors with graphene is possible because of the higher speed at which electrons in graphene move compared to electrons in silicon. Researchers are also developing lithography techniques that can be used to fabricate integrated circuits based on graphene.

Sensors to diagnose diseases. These sensors are based upon graphene’s large surface area and the fact that molecules that are sensitive to particular diseases can attach to the carbon atoms in graphene. For example, researchers have found that graphene, strands of DNA, and fluorescent molecules can be combined to diagnose diseases. A sensor is formed by attaching fluorescent molecules to single strand DNA and then attaching the DNA to graphene.  When an identical single strand DNA combines with the strand on the graphene a double strand DNA if formed that floats off from the graphene, increasing the fluorescence level. This method results in a sensor that can detect the same DNA for a particular disease in a sample.

Membranes for more efficient separation of gases. These membranes are made from sheets ofgraphene in which nanoscale pores have been created. Because graphene is only one atom thick researchers believe that gas separation will require less energy than thicker membranes.

Chemical sensors effective at detecting explosives. These sensors contain sheets of graphene in the form of a foam which changes resistance when low levels of vapors from chemicals, such as ammonia, is present.

Graphene Company Directory

Graphene Company Product
Angstron Materials Graphene Supplier
Bluestone Global Tech Graphene Supplier
CrayoNano Semiconductor nanowires grown on graphene

 

Nanomaterial Applications using Nanocomposites

Applications being developed for nanocomposites include a nanotube-polymer nanocomposite to form a scaffold which speeds up replacement of broken bones, making a graphene-epoxy nanocomposite with very high strenght-to-weight ratios, a nanocomposite made from cellulous and nanotubes used to make a flexible battery.

 

nanocomposite is a matrix to which nanoparticles have been added to improve a particular property of the material. The properties of nanocomposites have caused researchers and companies to consider using this material in several fields.

A survey of  the applications of nanocomposites:

The following survey of nanocomposite applications introduces you to many of the uses being explored, including:

Producing batteries with greater power output. Researchers have developed a method to make anodes for lithium ion batteries from a composite formed with silicon nanospheres and carbon nanoparticles. The anodes made of the silicon-carbon nanocomposite make closer contact with the lithium electrolyte, which allows faster charging or discharging of power.

Speeding up the healing process for broken bones. Researchers have shown that growth of replacement bone is speeded up when a nanotube-polymer nanocomposite is placed as a kind of scaffold which guides growth of replacement bone. The researchers are conducting studies to better understand how this nanocomposite increases bone growth.

Producing structural components with a high strength-to-weight ratio.  For example an epoxy containing carbon nanotubes can be used to produce nanotube-polymer composite windmill blades. This results in a strong but lightweight blade, which makes longer windmill blades practical. These longer blades increase the amount of electricity generated by each windmill.

Using graphene to make composites with even higher strength-to-weight ratios. Researchers have found that adding graphene to epoxy composites may result in stronger/stiffer components than epoxy composites using a similar weight of carbon

nanotubes. Graphene appears to bond better to the polymers in the epoxy, allowing a more effective coupling of the graphene into the structure of the composite. This property could result in the manufacture of components with higher strength-to-weight ratios for such uses as windmill blades or aircraft components.

Making lightweight sensors with nanocomposites. A polymer-nanotube nanocomposite conducts electricity; how well it conducts depends upon the spacing of the nanotubes. This property allows patches of polymer-nanotube nanocomposite to act as stress sensors on windmill blades. When strong wind gusts bend the blades the nanocomposite will also bend. Bending changes the nanocomposite sensor’s electrical conductance, causing an alarm to be sounded. This alarm would allow the windmill to be shut down before excessive damage occurs.

Using nanocomposites to make flexible batteries. A nanocomposite of cellulous materials and nanotubes could be used to make a conductive paper. When this conductive paper is soaked in an electrolyte, a flexible battery is formed.

Making tumors easier to see and remove. Researchers are attempting to join magnetic nanoparticles and fluorescent nanoparticles in a nanocomposite particle that is both magnetic and fluorescent. The magnetic property of the nanocomposite particle makes the tumor more visible during an MRI procedure  done prior to surgery. The fluorescent property of the nanocomposite particle could help the surgeon to better see the tumor while operating.

Nanocomposite Company Directory

Company Product
Nanosonic Metal Rubber™ nanocomposites
InMat Nanocomposite coatings
Nanocyl EPOCYL™ epoxy resins reinforced with carbon nanotubes
MesaCoat Nanocomposite coatings
NanoComposites Nanocomposite materials

More Nanocomposite Companies

 

Nanomaterial Applications using Nanofibers

Applications being developed for nanofibers include stimulating the production of cartilage in damaged joints, piezoelectric nanofibers that can be woven into clothing to produce electricty for cell phones or other devices, carbon nanofibers that can improve the preformance flame retandant in funiture.

 

A nanofiber is a fiber with a diameter of 100 nanometers or less. The properties of nanofibers have caused researchers and companies to consider using this material in several fields.

A survey of the applications of nanofibers:

Researchers are using nanofibers to capture individual cancer cells circulating in the blood stream. They use nanofibers coated with antibodies that bind to cancer cells, trapping the cancer cell for analysis.

Nanofibers can stimulate the production of cartilage in damaged joints. Three different approaches to the use of nanofibers to stimulate cartilage are being taken by researchers at John Hopkins University, at Northwestern University and at the University of Pennsylvania.

Reseachers are using nanofibers to delivery thrapeutic drugs. The have developed an elastic material that is embedded with needle like carbon nanofibers. The material is intended to be used as balloons which are inserted next diseased tissue, and then inflated. When the balloon is inflated the carbon nanofibers penetrate diseased cells and delivery therapeutic drugs.

Researchers at MIT have used carbon nanofibers to make lithium ion battery electrodes that show four times the storage capacity of current lithium ion batteries.

The next step beyond lithium-ion batteries may be lithium sulfur batteries (the cathode contains the sulfur), which have the capability of storing several times the energy of lithium-ion  batteries. Researchers at Stanford University are using cathodes made up of carbon nanofibers encapsulating the sulfur.

Researchers are using nanofibers to make sensors that change color as they absorb chemical vapors. They plan to use these sensors to show when the absorbing material in a gas mask becomes saturated.

Researchers have developed piezoelectric nanofibers that are flexible enough to be woven into clothing. The fibers can turn normal motion into electricity to power your cell phone and other mobile electronic devices.

Flame retardant formed by coating the foam used in furniture with carbon nanofibers.

 

Nanomaterial Applications using Nanoparticles

Applications being developed for nanoparticles include deliver chemotherapy drugs directly to cancer tumors, resetting the immune system to prevent autoimmune diseases, delivering drugs to damaged regions of arteries to fight cardiovascular disease, create photocatalysts that produce hydrogen from water, reduce the cost of producing fuel cells and solar cells, clean up oil spills, water pollution and air pollution

 

Nanoparticles have one dimension that measures 100 nanometers or less. The properties of many conventional materials change when formed from nanoparticles. This is typically because nanoparticles have a greater surface area per weight than larger particles which causes them to be more reactive to some other molecules.

Nanoparticles are used, or being evaluated for use, in many fields. The list below introduces several of the uses under development.

Nanoparticle Applications in Medicine

The use of polymeric micelle nanoparticles to deliver drugs to tumors.

The use of polymer coated iron oxide nanoparticles to break up clusters of bacteria, possibly allowing more effective treatment of chronic bacterial infections.

The surface change of protein filled nanoparticles has been shown to affect the ability of the nanoparticle to stimulate immune responses. Researchers are thinking that these nanoparticles may be used in inhalable vaccines.

Researchers at Rice University have demonstrated that cerium oxide nanoparticles act as an antioxidant to remove oxygen free radicals that are present in a patient’s bloodstream following a traumatic injury. The nanoparticles absorb the oxygen free radicals and then release the oxygen in a less dangerous state, freeing up the nanoparticle to absorb more free radicals.

Researhers are developing ways to use carbon nanoparticles called nanodiamonds in medical applications. For example nanodiamonds with protein molecules attached can be used to increase bone growth around dental or joint implants.

Researchers are testing the use of chemotherapy drugs attached to nanodiamonds to treat brain tumors. Other researchers are testing the use of chemotherapy drugs attached to nanodiamonds to treat leukemia.

More about Nanotechnology in Medicine

Nanoparticle Applications in Manufacturing and Materials

Ceramic silicon carbide nanoparticles dispersed in magnesium produce a strong, lightweight material.

A synthetic skin, that may be used in prosthetics, has been demonstrated with both self healing capability and the ability to sense pressure. The material is a composite of nickel nanoparticles and a polymer. If the material is held together after a cut it seals together in about 30 minutes giving it a self healing ability. Also the electrical resistance of the material changes with pressure, giving it a sense ability like touch.

Silicate nanoparticles can be used to provide a barrier to gasses (for example oxygen), or moisture in a plastic film used for packaging. This could slow down the process of spoiling or drying out in food.

Zinc oxide nanoparticles can be dispersed in industrial coatings to protect wood, plastic, and textiles from exposure to UV rays.

Silicon dioxide crystalline nanoparticles can be used to fill gaps between carbon fibers, thereby strengthening tennis racquets.

Silver nanoparticles in fabric are used to kill bacteria, making clothing odor-resistant.

Nanoparticle Applications and the Environment

Researchers are using photocatalytic copper tungsten oxide nanoparticles to break down oil into biodegradable compounds. The nanoparticles are in a grid that provides high surface area for the reaction, is activated by sunlight and can work in water, making them useful for cleaning up oil spills.

Researchers are using gold nanoparticles embedded in a porous manganese oxide as a room temperature catalyst to breakdown volatile organic pollutants in air.

Iron nanoparticles are being used to clean up carbon tetrachloride pollution in ground water.

Iron oxide nanoparticles are being used to clean arsenic from water wells.

Nanoparticle Applications in Energy and Electronics

Researchers have used nanoparticles called nanotetrapods studded with nanoparticles of carbon to develop low cost electrodes for fuel cells. This electrode may be able to replace the expensive platinum needed for fuel cell catalysts.

 

Researchers at Georgia Tech, the University of Tokyo and Microsoft Research have developed a method to print prototype circuit boards using standard inkjet printers. Silver nanoparticle ink was used to form the conductive lines needed in circuit boards.

Combining gold nanoparticles with organic molecules creates a transistor known as a NOMFET (Nanoparticle Organic Memory Field-Effect Transistor). This transistor is unusual in that it can function  in a way similar to synapses in the nervous system.

catalyst using platinum-cobalt nanoparticles is being developed for fuel cells that produces twelve times more catalytic activity than pure platinum. In order to achieve this performance, researchers anneal nanoparticles to form them into a crystalline lattice, reducing the spacing between platinum atoms on the surface and increasing their reactivity.

Researchers have demonstrated that sunlight, concentrated on nanoparticles, can produce steam with high energy efficiency. The “solar steam device” is intended to be used in areas of developing countries without electricity for applications such as purifying water or disinfecting dental instruments.

A lead free solder reliable enough for space missions and other high stress environments using copper nanoparticles.

Silicon nanoparticles coating anodes of lithium-ion batteries can increase battery power and reduce recharge time.

Semiconductor nanoparticles are being applied in a low temperature printing process that enables the  manufacture of low cost solar cells.

A layer of closely spaced palladium nanoparticles is being used in a hydrogen sensor. Whenhydrogen is absorbed, the palladium nanoparticles swell, causing shorts between nanoparticles. These shorts lower the resistance of the palladium layer.

Nanoparticle Company Directory

Company Products
CytImmune Gold nanoparticles for targeted delivery of drugs to tumors
Invitrogen Qdots for medical imaging
Antaria Zinc oxide nanoparticles used in coatings to reduce UV exposure
Nanoledge Epoxy resins strengthened with nanoparticles

 

Nanomaterial Applications using Nanowires

Applications being developed for carbon nanotubes include using zinc oxide nanowires in a flexible solar cell, silver chloride nanowires to decompose organic molecules in polluted water, using nanowires made from iron and nickel to make dense computer memory – called “race track memory

 

The properties of nanowires have caused researchers and companies to consider using this material in several fields.

Nanowires Applications in Energy

Researchers at MIT have developed a solar cell using graphene coated with zinc oxide nanowires. The researchers believe that this method will allow the production of low cost flexible solar cells at high enough efficiency to be competive.

Sensors powered by electricity generated by piezoelectric zinc oxide nanowires. This could allowsmall, self contained, sensors powered by mechanical energy such as tides or wind

Researchers are using a method called Aerotaxy to grow semiconducting nanowires on gold nanoparticles. They plan to use self assembly techniques to align the nanowires on a substrate; forming a solar cell or other electrical devices. The gold nanoparticles replace the silicon substrate on which conventional semiconductor based solar cells are built.

Researchers at the Nies Bohr Institute have determined that sunlight can be concentrated in nanowires due to a resonance effect. This effect can result in more efficient solar cells, allowing more of the energy from the sun to be converted to electricity.

Using light absorbing nanowires embedded in a flexible polymer film is another method being developed to produce low cost flexible solar panels.

Researchers at Lawrence Berkeley have demonstrated an inexpensive process for making solar cells. These solar cells are composed of cadmium sulfide nanowires coated with copper sulfide.

Researchers at Stanford University have grown silicon nanowires on a stainless steel substrate and demonstrated that batteries using these anodes could have up to 10 times the power density of conventional lithium ion batteries. Using silicon nanowires, instead of bulk silicon fixes a problem of the silicon cracking, that has been seen on electrodes using bulk silicon. The cracking is caused because the silicon swells it absorbs lithium ions while being recharged, and contracts as the battery is discharged and the lithium ions leave the silicon. However the researchers found that while the silicon nanowires swell as lithium ions are absorbed during discharge of the battery and contract as the lithium ions leave during recharge of the battery the nanowires do not crack, unlike anodes that used bulk silicon.

Nanowire Applications in the Enviroment

Using silver chloride nanowires as a photocatalysis to decompose organic molecules in polluted water.

Using an electrified filter composed of silver nanowires, carbon nanotubes and cotton to kill bacteria in water.

Using nanowire mats to absorb oil spill

Nanowire Applications in Electronics

Using electrodes made from nanowires that would enable flat panel displays to be flexible as well as thinner than current flat panel displays.

Using nanowires to build transistors without p-n junctions.

Using nanowires made of an alloy of iron and nickel to create dense memory devices. By applying a current magnetized sections along the length of the wire. As the magnetized sections move along the wire, the data is read by a stationary sensor. This method is called race track memory.

Using silver nanowires embedded in a polymer to make conductive layers that can flex, without damaging the conductor.

Sensors using zinc oxide nano-wire detection elements capable of detecting

M.Tech U III The effect of chemistry of nanostructures

Unit-III  The effect of chemistry of nanostructures: Modification of nanoparticles, Langmuir Blodgett films, Self assembled surface films, Binding of molecules on solid substrate surfaces, Molecular nanostructures, Strategies of molecular construction, Synthetic supramolecules.

 

Modification of nanoparticles

Nanoparticles and nanocomposites are used in a wide range of applications in various fields, such as medicine, textiles, cosmetics, agriculture, optics, food packaging, optoelectronic devices, semiconductor devices, aerospace, construction and catalysis. Nanoparticles can be incorporated into polymeric nanocomposites. Polymeric nanocomposites consisting of inorganic nanoparticles and organic polymers represent a new class of materials that exhibit improved performance compared to their microparticle counterparts. It is therefore expected that they will advance the field of engineering applications. Incorporation of inorganic nanoparticles into a polymer matrix can significantly affect the properties of the matrix. The resulting composite might exhibit improved thermal, mechanical, rheological, electrical, catalytic, fire retardancy and optical properties. The properties of polymer composites depend on the type of nanoparticles that are incorporated, their size and shape, their concentration and their interactions with the polymer matrix. The main problem with polymer nanocomposites is the prevention of particle aggregation. It is difficult to produce monodispersed nanoparticles in a polymer matrix because nanoparticles agglomerate due to their specific surface area and volume effects. This problem can be overcome by modification of the surface of the inorganic particles. The modification improves the interfacial interactions between the inorganic particles and the polymer matrix. There are two ways to modify the surface of inorganic particles. The first is accomplished through surface absorption or reaction with small molecules, such as silane coupling agents, and the second method is based on grafting polymeric molecules through covalent bonding to the hydroxyl groups existing on the particles. The advantage of the second procedure over the first lies in the fact that the polymer-grafted particles can be designed with the desired properties through a proper selection of the species of the grafting monomers and the choice of grafting conditions.

 

Langmuir, Langmuir-Blodgett, Langmuir-Schaefer Technique

The Langmuir (L), Langmuir-Blodgett (LB) and Langmuir-Schaefer (LS) techniques enable fabrication and characterization of single molecule thick films with control over the packing density of molecules. They also enable the creation of multilayer structures with varying layer composition.

Langmuir, Langmuir-Blodgett, Langmuir-Schaefer—what is the difference?

When a monolayer is fabricated at the gas-liquid or liquid-liquid interface, the film is named Langmuir film. A Langmuir film can be deposited on a solid surface and is thereafter called Langmuir-Blodgett film (in the case of vertical deposition) or Langmuir-Schaefer film (in the case of horizontal deposition). Langmuir-Schaefer is often seen just as a variant of Langmuir-Blodgett deposition.

Langmuir film, Langmuir-Blodgett deposition, Langmuir-Schaefer deposition and multilayers obtained after repeated deposition.

Langmuir Troughs (or Langmuir film balance) are used for Langmuir film fabrication and characterization. Langmuir-Blodgett troughs are used for Langmuir-Blodgett or Langmuir-Schaefer deposition. All KSV NIMA Troughs are modular and when equipped with the right modules can be used for Langmuir film fabrication or characterization as well as Langmuir-Blodgett and Langmuir-Schaefer deposition.

The components of L and LB Troughs

Langmuir troughs include a set of barriers (2), a Langmuir trough top (3*) and a surface pressure sensor (4) as standard. The software-controlled barriers are placed at the interface and compress the monolayer. The trough top holds the liquid phase where monolayers are fabricated. The trough top is often made of hydrophobic material that improves sub-phase containment. The surface pressure sensor provides information about monolayer packing density.

Langmuir-Blodgett troughs include a set of barriers (2), a Langmuir-Blodgett trough top (3*), a surface pressure sensor (4) and a dipping mechanism (5) as standard. The Langmuir-Blodgett trough top holds the liquid phase and has a well in the center to allow space for solid substrate dipping through the monolayer. The dipping mechanism holds the solid substrate and enables controlled deposition cycle(s).

Please note that for Langmuir-Schaefer deposition, the Langmuir-Blodgett trough top is not always necessary and can in some cases be replaced by a Langmuir trough top.

KSV NIMA L & LB Trough modules

  1. Frame
  2. Barriers
  3. Trough top
  4. Surface pressure sensor
  5. Dipping mechanism (LB option)
  6. Interface unit

KSV NIMA troughs are built on a frame (1) that enables outstanding modularity; a Langmuir-Blodgett trough top can be easily switched with a Langmuir trough top. The dipping mechanism can also be added or removed for simple conversion between Langmuir and Langmuir-Blodgett configurations. All KSV NIMA troughs come with an interface unit (6) that controls the instrument and displays key measurements.

Langmuir film fabrication

Prepare the amphiphile molecules that will create a monolayer in a water insoluble solvent. The sub-phase, typically water, is held in the hydrophobic trough top that gives good sub-phase containment. When the amphiphile solution is deposited on the water surface with a microsyringe, the solution spreads rapidly to cover the available area.  As the solvent evaporates, a monolayer forms at the air-water interface and a Langmuir film is created.
The software-controlled barriers located at the interface then compress the monolayer until the surface pressure sensor indicates maximum packing density.

A compressed, monolayer film can be considered as a two-dimensional solid with a surface area to volume ratio far above that of bulk materials. In these conditions, materials often yield fascinating new properties. Experimentation using Langmuir troughs enables inference and understanding about how particular molecules pack when confined in two dimensions. The surface pressure-area isotherm can also provide a measure of the average area per molecule and the compressibility of the monolayer.

Surface pressure—area isotherms of a Langmuir film and molecules in different phases.

Langmuir film characterization

Langmuir films fabricated in a Langmuir trough can be studied by analyzing surface pressure isotherms, isochors, and other data measured with the trough or with a complementary characterization instrument.

KSV NIMA Langmuir troughs enable measurements of:

Measurement

Information

Isotherms
Structure, area, interactions, phase transitions, compressibility, hysteresis
Isobar/Isochors
Stability
Surface potential*
Dissociation, orientation, interactions
Dilational rheology
Film viscoelastic properties
Kinetics
Polymerization and enzyme kinetics
Conductivity
Lateral conductivity
Environmental monitoring
pH* and temperature

*Optional

KSV NIMA Microscopy Troughs are special troughs equipped with a sapphire window in the top. The sapphire window allows high optical transmission down to 200 nm, which is suitable for visible light or UV microscopy. Troughs suitable for both upright and inverted microscopes are available.

For more information about Langmuir film microscopy, see:

Popular complementary characterization techniques include: Brewster Angle Microscopy (for film visualization), FTIR spectrometry such as PM-IRRAS (for determination of orientation and chemical composition), Interfacial Shear Rheometry (for viscoelastic properties), Surface Potential Sensing (for determination of changes in packing and orientation), Vibrational spectroscopy, UV-VIS absorbance spectroscopy, and X-ray reflectometry.

For more information, see:

Langmuir-Blodgett film deposition

Langmuir films can be transferred to solid surfaces with preserved density, thickness and homogeneity of the sample. This allows the assembly of organized multilayer structures with varying layer compositions. Compared to other organic thin film deposition techniques, LB is less limited by the molecular structure of the functional molecule and is often the only technique that can be used for bottom-up assembly.

LB deposition is traditionally carried out in the ‘solid’ phase where surface pressure is high enough to ensure sufficient cohesion in the monolayer. This means that attraction between the molecules in the monolayer is sufficient to prevent the monolayer from falling apart during transfer to the solid substrate and ensures the build up of homogeneous multilayers. The surface pressure that gives the best results depends on the nature of the monolayer and is usually established empirically. Generally, amphiphiles can seldom be successfully deposited at surface pressures lower than 10 mN/m, and at surface pressures above 40 mN/m collapse and film rigidity often pose problems. When the solid substrate is hydrophilic (glass, SiO2 etc.) the first layer is deposited by raising the solid substrate from the sub-phase through the monolayer, whereas if the solid substrate is hydrophobic (HOPG, silanized SiO2 etc.) the first layer is deposited by lowering the substrate into the sub-phase through the monolayer.

Monolayers can be held at a constant surface pressure by a computer-controlled feedback between the surface pressure sensor and the compressing barriers. This is useful when producing LB films to guarantee the homogeneity of the film deposited.

In the case of Langmuir-Blodgett (LB) deposition, the solid substrate is dipped through the Langmuir film and extra space is required below the monolayer. It means the Langmuir film has to be fabricated with a LB-trough top with a sufficient well size for the substrate. The dipping mechanism holds the solid substrate and enables controlled deposition cycle(s). The Langmuir-Schaefer (LS) technique can be performed with a Langmuir trough top, as no additional depth is required below the monolayer.

Repeated deposition can be achieved to obtain well-organized multilayers on the solid substrate. LB and LS cycles can also be combined to obtain desired structures and thicknesses. The most common multilayer deposition is the Y-type multilayer, which is produced when the monolayer deposits to the solid substrate in both up and down directions. When the monolayer deposits only in the up or down direction the multilayer structure is called either Z-type or X-type. Intermediate structures are sometimes observed for some LB multilayers and they are often referred to be XY-type multilayers.

Various LB deposition possibilities on hydrophobic and hydrophilic substrates.

Some special LB deposition troughs such as the KSV NIMA Alternate-Layer Langmuir-Blodgett Deposition Trough are designed for fully automatic LB multi-deposition from two different Langmuir films.

Alternate LB deposition with the KSV NIMA LB Trough Alternate

There are several parameters that affect the type of LB film produced. These include: the nature of the spread film, the sub-phase composition and temperature, the surface pressure during the deposition and the deposition speed, the type and nature of the solid substrate and the time the solid substrate is stored in air or in the sub-phase between the deposition cycles. The quantity and the quality of the deposited monolayer on a solid support are measured by the transfer ratio (t.r.). This is defined as the ratio between the decrease in monolayer area during a deposition stroke, Al, and the area of the substrate, As. An ideal transfer has a t.r. that is equal to 1.

Langmuir-Blodgett film characterization



Many properties of LB films depend on the properties of the Langmuir film it was created from. LB films can be characterized for additional information and checked for the quality of the deposition. Commonly used techniques are include: PM-IRRAS (FTIR spectrometry), Surface Plasmon Resonance, Quartz Cristal Microbalance, Ellipsometry, Vibrational spectroscopy, UV-VIS absorbance spectroscopy, X-ray reflectometry etc.

Self-Assembled Monolayers

Self-assembled monolayers (SAMs) are ordered molecular assemblies formed by the adsorption of amphiphilic, surfactant-type molecules on solid surfaces. The substrate is generally immersed into a dilute solution of the film molecules or suitable precursors thereof and a monolayer film forms spontanously in a time span of a few minutes to a few hours.

A typical amphiphilic molecule (octadecyltrichlorosilane), consisting of a long-chain alkyl group (C18H37) and a polar head group (SiCl3), which forms SAMs on various oxidic substrates.

The driving force for the surface aggregation and self-assembly is i) a covalent bond formation of the film molecules with the substrate surface via suitable functional groups and ii) intermolecular, van der Waals-type interactions between the hydrocarbon chains of the film molecules.

Formation of Self-Assembled Monolayers

According to the type of film-substrate bonding, SAMs can be grouped into the following categories:

  • Organosulfur compounds (thiols, thioethers, disulfides etc.) on late transition metals (gold, silver, copper, mercury)
  • Organosilicon compounds (alkylchlorosilanes, alkylalkoxysilanes) on metal and nonmetal oxides (Al2O3, TiO2, SnO2, SiO2, glass)
  • Fatty acids on metal oxides (AgO, CuO, Al2O3)
  • Alkylphosphonic acids on metal and nonmetal surfaces primed with coordinating transition metal ions (Zr4+, Hf4+, Cu2+, Zn2+)

M.Tech U-II Building Blocks of Nanotechnology

Unit-II  Building Blocks of Nanotechnology: covalent architecture, coordinated architecture and weakly bound aggregates, Interactions and topology,Chemical Properties: The effect of nanoscale metals on chemical reactivity, Effect of nanostructure on mass transport, Metal nanocrystallites , Supported nanoscale catalysts.

Building Blocks of Nanotechology

A buckyball. Source: Office of Basic Energy Science/U.S Dept. of Energy.

Nanotechnology is a field that’s just being established, and although there are big plans for the smallest of technologies, right now, most of what nanotechnologists have accomplished falls into three categories: new materials—usually chemicals—made by assembling atoms in new ways; new tools to make those materials; and the beginnings of tiny molecular machines.

Richard E. Smalley, winner of the 1996 Nobel Prize in Chemistry for the discovery of a structure of carbon atoms known as a “buckyball.” (Image Source: Brookhaven National Laboratory)

Some of the primary building blocks in nanotechnology are buckminsterfullerenes (almost always known as buckyballs or fullerenes), which are clumps of molecules that look like soccer balls. In 1984Richard Smalley, Robert Curl, and Harold Kroto were investigating an amazing molecule consisting of 60 linked atoms of carbon. Smalley worked these atoms into shapes he called “fullerenes,” a name based on architect Buckminster Fuller’s “geodesic” domes of the 1930s and first suggested by Japan’s Eiji Osawa. Sumio Iijima, Smalley, and others found similar structures in the form of tubes, and found that fullerenes had unique chemical and electrical properties. Fullerenes became nanotech’s first major new material. But what to do with them? Engineers turned their attention to finding some practical use for these interesting molecules.

The letters “IBM” spelled in xenon atoms, as imaged by the atomic force microscope. Courtesy: IBM.

While engineers thought about practical uses for fullerenes another discovery in search of an application was being made. In 1981 Gerd Karl Binnig and Heinrich Rohrer invented the scanning tunneling microscope or STM, which has a tiny tip so sensitive that it can in effect “feel” the surface of a single atom. It then sends information about the surface to a computer that reconstructs an image of the atomic surface on a display screen. If that weren’t amazing enough, a little later, researchers discovered that the tip of the STM could actually move atoms around, and Donald Eigler and a team at IBM staged a dramatic demonstration of this new ability, spelling out “IBM”. Researchers believed they had a tool, the atomic force microscope (AFM), that could build things atom-by-atom. But, like the discovery of fullerenes, it remained to be seen if anything useful could actually be built this way.

The development of tools such as AFMs coincided with the introduction of very powerful new computers and software that scientists could use to simulate and visualize chemical reactions or “build” virtual atoms and molecules. This was especially useful for scientists working with complex chemical molecules, particularly DNA. Researchers recognized that the actions of DNA resembled some of the things nanotechnologists were now calling for—the use of molecules to construct other molecules, the self-replication of molecules, and the use of molecule-size mechanical devices. Perhaps DNA (or its cousin, RNA) could be modified to create the first nanomachines?

Ned Seeman has succeeded in manipulating strands of DNA into customized molecules with multiple interconnections. He believes this is the first step in doing more complex things with DNA, such as using it to create molecular machinery. Source: NYU.

Geneticists had already found ways to use DNA taken from bacteria to make a nano-scale replicator used for scientific research. By modifying some of the chemical reactions that take place in natural DNA, genetic engineers had figured out a way to make copies of nearly any DNA molecule they wanted to study. But with the computers and tools available to them by the 1990s, they began using DNA or DNA-like molecules to do other things—like construct new chemicals or tiny machines. Many researchers began investigating ways to make proteins—the components from which DNA is made—that would perform useful tasks, such as interacting with other materials or living cells to create new materials or perhaps attack diseases. One of the first breakthroughs was Professor Nadrian Seeman’s demonstration of a tiny “robot arm” made from modified DNA. While the arm could not yet really do anything useful, it did demonstrate the concept.

Building Blocks of Nanotechnology Microgyroscope.jpg

Meanwhile, electronics researchers approached nanotechnology from another direction. Since 1959, engineers had etched and coated silicon chips using a variety of processes to make integrated circuits (ICs). The transistors and other chip elements reached nano-scale in the late 1990s. They also used these same techniques to develop the first micromachines—microscopic devices with actual moving parts. Some of the early versions of these were simply intended to demonstrate the process without doing anything particularly useful, such as a tiny guitar with a string that could be plucked using an atomic force microscope. But in the late 1980s these began to be commercialized as machines-on-a-chip, or micro-electrco-mechanical systems (MEMs), which combine ICs and tiny mechanical elements. However useful MEMs are, most engineers feel that the techniques used to make ordinary ICs will never be refined enough to make true nanotechnologies. For that reason, engineers are now concentrating on discovering entirely new ways to make ICs, building them from the ground up rather than cutting and etching “bulk” silicon slices.

With the appearance of protein-based chemistry and other techniques in the 1990s, researchers began looking both for practical uses for nanotechnology and new ways to make nano-molecules or micromachines. A different but related problem was that of making nanomolecules in large numbers. A single nanomachine or nanocircuit for example, would not be able to do enough work to make a difference in the real world—thousands or millions might be needed. Engineers needed ways to turn out their nanomachines in huge numbers, and so they began looking for a way to make a nano-scale machine or molecule that would assemble other nano-scale machines or molecules. K. Eric Drexler called it a “self assembler,” and scientists believe that it will be one of the keys to making certain kinds of nanotechnology useful and practical. To date, very few practical nanotechnologies and no self-assemblers have been used outside the laboratory.

Synthesis

Although it seems at first that Nature has provided a limited number of basic building blocks-amino acids, lipids, and nucleic acids-the chemical diversity of these molecules and the different ways they can be polymerized or assembled provide an enormous range of possible structures. Furthermore, advances in chemical synthesis and biotechnology enable one to combine these building blocks, almost at will, to produce new materials and structures that have not yet been made in Nature. These self-assembled materials often have enhanced properties as well as unique applications.

The selected examples below show ways in which clever synthetic methodologies are being harnessed to provide novel biological building blocks for nanotechnology.

The protein polymers produced by Tirrell and coworkers (1994) are examples of this new methodology. In one set of experiments, proteins were

Figure 7.2
Top: a 36-mer protein polymer with the repeat sequence (ulanine-glycine)3 – glutamic acid – glycine. Bottom: idealized folding of this protein polymer, where the glutamic acid sidechains (+) are on the surface of the folds.

designed from first principles to have folds in specific locations and surface-reactive groups in other places (Figure 7.2) (Krejchi et al. 1994; 1997). One of the target sequences was -((AG)3EG)36. The hypothesis was that the AG regions would form hydrogen-bonded networks of beta sheets and that the glutamic acid would provide a functional group for surface modification. Synthetic DNAs coding for these proteins were produced, inserted into an E. coli expression system, and the desired proteins were produced and harvested. These biopolymers formed chain-folded lamellar crystals with the anticipated folds. In addition to serving as a source of totally new materials, this type of research also enables us to test our understanding of amino acid interactions and our ability to predict chain folding.

Biopolymers produced via biotechnology are monodisperse; that is, they have precisely defined and controlled chain lengths; on the other hand, it is virtually impossible to produce a monodisperse synthetic polymer. It has recently been shown that polymers with well-defined chain lengths can have unusual liquid crystalline properties. For example, Yu et al. (1997) have shown that bacterial methods for polymer synthesis can be used to produce poly(gamma-benzyl alpha L-glutamate) that exhibits smectic ordering in solution and in films. The distribution in chain length normally found for synthetic polymers makes it unusual to find them in smectic phases. This work is important in that it suggests that we now have a route to new smectic phases whose layer spacings can be controlled on the scale of tens of nanometers.

The biotechnology-based synthetic approaches described above generally require that the final product be made from the natural, or L-amino acids. Progress is now being made so that biological machinery (e.g., E. coli), can be co-opted to incorporate non-natural amino acids such as b -alanine or dehydroproline or fluorotyrosine, or ones with alkene or alkyne functionality (Deming et al. 1997). Research along these lines opens new avenues for producing controlled-length polymers with controllable surface properties, as well as biosynthetic polymers that demonstrate electrical phenomena like conductivity. Such molecules could be used in nanotechnology applications.

Novel chemical synthesis methods are also being developed to produce “chimeric” molecules that contain organic turn units and hydrogen-bonding networks of amino acids (Winningham and Sogah 1997). Another approach includes incorporating all tools of chemistry into the synthesis of proteins, making it possible to produce, for example, mirror-image proteins. These proteins, by virtue of their D-amino acid composition, resist biodegradation and could have important pharmaceutical applications (Muir et al. 1997).

Arnold and coworkers are using a totally different approach to produce proteins with enhanced properties such as catalytic activity or binding affinity. Called “directed evolution,” this method uses random mutagenesis and multiple generations to produce new proteins with enhanced properties. Directed evolution, which involves DNA shuffling, has been used to obtain esterases with five- to six-fold enhanced activity against p-nitrobenzyl esters (Moore et al. 1997).

Assembly

The ability of biological molecules to undergo highly controlled and hierarchical assembly makes them ideal for applications in nanotechnology. The self-assembly hierarchy of biological materials begins with monomer molecules (e.g., nucleotides and nucleosides, amino acids, lipids), which form polymers (e.g., DNA, RNA, proteins, polysaccharides), then assemblies (e.g., membranes, organelles), and finally cells, organs, organisms, and even populations (Rousseau and Jelinski 1991, 571-608). Consequently, biological materials assembly on a very broad range of organizational length scales, and in both hierarchical and nested manners (Aksay et al. 1996; Aksay 1998). Research frontiers that exploit the capacity of biomolecules and cellular systems to undergo self-assembly have been identified in two recent National Research Council reports (NRC 1994 and 1996). Examples of self-assembled systems include monolayers and multilayers, biocompatible layers, decorated membranes, organized structures such as microtubules and biomineralization, and the intracellular assembly of CdSe semiconductors and chains of magnetite.

A number of researchers have been exploiting the predictable base-pairing of DNA to build molecular-sized, complex, three-dimensional objects. For example, Seeman and coworkers (Seeman 1998) have been investigating these properties of DNA molecules with the goal of forming complex 2-D and 3-D periodic structures with defined topologies. DNA is ideal for building molecular nanotechnology objects, as it offers synthetic control, predictability of interactions, and well-controlled “sticky ends” that assemble in highly specific fashion. Furthermore, the existence of stable branched DNA molecules permits complex and interlocking shapes to be formed. Using such technology, a number of topologies have been prepared, including cubes (Chen and Seeman 1991), truncated octahedra (Figure 7.3) (Zhang and Seeman 1994), and Borromean rings (Mao et al. 1997).

Other researchers are using the capacity of DNA to self-organize to develop photonic array devices and other molecular photonic components (Sosnowski et al. 1997). This approach uses DNA-derived structures and a microelectronic template device that produces controlled electric fields. The electric fields regulate transport, hybridization, and denaturation of oligonucleotides. Because these electric fields direct the assembly and transport of the devices on the template surface, this method offers a versatile way to control assembly.

There is a large body of literature on the self-assembly on monolayers of lipid and lipid-like molecules (Allara 1996, 97-102; Bishop and Nuzzo 1996). Devices using self-assembled monolayers are now available for analyzing the binding of biological molecules, as well as for spatially tailoring the

Figure 7.3.
Idealized truncated octahedron assembled from DNA. This view is down the four-fold axis of the squares. Each edge of the octahedron contains two double-helical turns of DNA.

surface activity. The technology to make self-assembled monolayers (SAMs) is now so well developed that it should be possible to use them for complex electronic structures and molecular-scale devices.

Research stemming from the study of SAMs (e.g., alkylthiols and other biomembrane mimics on gold) led to the discovery of “stamping” (Figure 7.4) (Kumar and Whitesides 1993). This method, in which an elastomeric stamp is used for rapid pattern transfer, has now been driven to < 50 nanometer scales and extended to nonflat surfaces. It is also called “soft lithography” and offers exciting possibilities for producing devices with unusual shapes or geometries.

Self-assembled organic materials such as proteins and/or lipids can be used to form the scaffolding for the deposition of inorganic material to form ceramics such as hydroxyapatite, calcium carbonate, silicon dioxide, and iron oxide. Although the formation of special ceramics is bio-inspired, the organic material need not be of biological origin. An example is production of template-assisted nanostructured ceramic thin films (Aksay et al. 1996).

A particularly interesting example of bio-inspired self-assembly has been described in a recent article by Stupp and coworkers (Stupp et al. 1997). This work, in which organic “rod-coil” molecules were induced to self-assemble, is significant in that the molecules orient themselves and self-assemble over a wide range of length scales, including mushroom-shaped clusters (Figure 7.5); sheets of the clusters packed side-by-side; and thick films, where the sheets pack in a head-to-tail fashion. The interplay between hydrophobic and hydrophilic forces is thought to be partially responsible for the controlled assembly.

 

Molecular building blocks and development strategies for molecular nanotechnology

 

If we are to manufacture products with molecular precision, we must develop molecular manufacturing methods. There are basically two ways to assemble molecular parts: self assembly and positional assembly. Self assembly is now a large field with an extensive body of research. Positional assembly at the molecular scale is a much newer field which has less demonstrated capability, but which also has the potential to make a much wider range of products. There are many arrangements of atoms which seem either difficult or impossible to make using the methods of self assembly alone. By contrast, positional assembly at the molecular scale should make possible the synthesis of a much wider range of molecular structures.

One of the fundamental requirements for positional assembly of molecular machines is the availability of molecular parts. One class of molecular parts might be characterized as molecular building blocks, or MBBs. With an atom count ranging anywhere from ten to ten thousand (and even more), such MBBs would be synthesized and positioned using existing (or soon to be developed) methods. Thus, in contrast to investigations of the longer term possibilities of molecular manufacturing (which often rely on mechanisms and systems that are likely to take many years or even decades to develop), investigations of MBBs focus on nearer term developmental pathways.

Introduction

Making a self replicating diamondoid assembler able to manufacture a wide range of products is likely to require several major stages, as its direct manufacture using existing technology seems quite difficult (Drexler, 1992; Merkle, 1996). For example, existing proposals call for the use of highly reactive tools in a vacuum or noble gas environment (Merkle, 1997d; Musgrave et al. 1991; Sinnot et al. 1994; Brenner et al. 1996; Brenner 1990). This requires an extremely clean environment and very precise and reliable positional control (Merkle, 1993b, 1997c) of the reactive tools. While these should be available in the future, they are not available today. Self replication has also been proposed as an important way to achieve low cost (Merkle, 1992).

A more attractive approach as a target for near term experimental efforts is the use of molecular building blocks (MBBs) (Krummenacker, 1992; Merkle, 1999). Such building blocks would be made from dozens to thousands of atoms (or more). Such relatively large building blocks would reduce the positional accuracy required for their assembly. Linking groups less promiscuous than the radicals proposed for the synthesis of diamond would also reduce the rate of incorrect side reactions in the presence of contaminants. Because this approach uses positional assembly at the molecular scale, and because positional assembly of molecules was, until recently, not a possibility that had been considered seriously, there has been remarkably little research in this area. As a consequence, the present paper will concentrate on providing perspective on the possibilities, along with a few examples to elucidate the more general principles. Further research into MBBs should prove well worth the effort.

The proposal to use molecular building blocks raises the obvious question: what do they look like? In this paper we consider a number of ideas and research directions which could be pursued to develop a firmer answer to this question.

Polymers are made from monomers, and each monomer reacts with two other monomers to form a linear chain. Synthetic polymers include nylon, dacron, and a host of others. Natural polymers include proteins and RNA (Watson et al., 1987) which, if the sequence of monomers forming the polymer is selected carefully, will fold into desired three dimensional shapes. While it is possible to make structures this way (as evidenced by the remarkable range of proteins found in biological systems), it is not the most intuitive approach (the protein folding problem is notoriously difficult).

A second drawback of this approach is the relative lack of stiffness of the resulting structure. The correct three dimensional shape is usually formed when many weak bonds combine to give the desired conformation greater stability and lower energy than the alternatives. However, this desired structure can usually be disrupted by changes in temperature, pressure, solvent, dissolved ions, or relatively modest mechanical force.

These limitations, caused in large measure by the restriction to two linking groups per monomer, motivates our investigation of MBBs with three or more linking groups.

An excellent review of well characterized linear rigid-rod oligomers formed by a variety of methods (Schwab et al., 1999) provides examples of the best exceptions to the general rule that polymers are floppy, though even here the rigidity is variable. However, giant molecules or supramolecular assemblies composed from the shorter and stiffer rods, particularly if well cross braced, might well prove to be extremely useful in the synthesis of stiff three dimensional structures.

The virtues of positional assembly, strength and stiffness

Before continuing, we digress to discuss the reasons for one of the primary design objectives for MBBs: stiffness. Strength and stiffness are desirable qualities in both individual MBBs and in the structures built from them. Intuitively, building things from marshmallows is usually less desirable than building them from wood or steel. More specifically, we expect to use the intermediate systems we build from MBBs to make more advanced systems, including assemblers. The manufacturing techniques that have been proposed for advanced systems rely heavily on positional assembly (Drexler, 1992; Merkle, 1993b). Positional assembly, in its turn, depends on the ability to position molecular parts with high precision despite thermal noise (Merkle, 1997c). To do this requires stiff materials from which to make the positional devices that are needed for positional assembly. We can’t make good robotic arms from marshmallows, we need something better.

There are two ways to assemble parts: self assembly and positional assembly. Self assembly is widely used at the molecular scale, and we find many examples of its use in biology (Watson, 1987). Positional assembly is widely used at the size scale of humans, and we find many examples of its use in manufacturing. Our inability to use positional assembly at the molecular scale with the same flexibility that we use it at the human scale seriously limits the range of structures that we can make.

By way of example, suppose we tried to make radios using self assembly. We would take the parts of the radio and put them into a bag, shake the bag, and pull out an assembled radio. This is a hard way to make a radio and if we demanded that all manufacturing take place using this approach our modern civilization would not exist.

By the same token, the range of things that can be made if we restrict ourselves to self assembly is much smaller than the range of things that can be made if we permit ourselves to add positional assembly to the other methods at our disposal. That this rather obvious point has not been more rapidly and generally understood with respect to the synthesis of molecular scale objects stems from the fact that we have never before been able to do positional assembly at the molecular scale. The idea of making a molecular structure by positionally assembling molecular parts is unfamiliar and different. Yet this capability, which has been demonstrated in nascent form by experimental work using the SPM (Scanning Probe Microscope) (Jung et al., 1996; Drexler et al., 1991, chapter 4 for a basic introduction) is clearly going to revolutionize our ability to make molecular structures and molecular machines (Drexler, 1992; Feynman, 1960).

Positional assembly is done using positional devices. At the scale of human beings, the major problem in positional assembly is overcoming gravity. Parts will fall down in a heap unless they are held in place by some strong positional device. At the molecular scale, the major problem in positional assembly is overcoming thermal noise. Parts will wiggle and jiggle out of position unless they are held in place by some stiff positional device (Merkle, 1999).

The fundamental equation relating positional uncertainty, temperature and stiffness is:

s2 = kbT/ks

Where s is the mean error in position, kb is Boltzmann’s constant, T is the temperature in Kelvins, and ks is the “spring constant” of the restoring force (Drexler, 1992). If ks is 10 N/m, the positional uncertainty s at room temperature is ~0.02 nm (nanometers). This is accurate enough to permit alignment of molecular parts to within a fraction of an atomic diameter.

A stiffness of 10 N/m is readily achievable with existing SPMs, but stiffness scales adversely with size. As we shrink a robotic arm, it gets less and less stiff and more and more compliant, and less and less able to position a part accurately in the face of thermal noise. To keep it stiff we have to make it from stiff parts. This is the fundamental driving force behind our desire to keep the MBB stiff.

In summary: stiffness is a fundamental design objective because we want to use positional assembly on molecular parts despite the positional uncertainty caused by thermal noise. This objective permeates our MBB design considerations.

The advantages and characteristics of molecular building blocks

Nanotechnology seeks the ability to make most structures consistent with physical law. When we use building blocks, particularly large building blocks, we drastically reduce the range of possible structures that we can make. If we adopt blocks of packed snow as building blocks we can make igloos, but we can’t make houses out of wood, steel, concrete or other building materials. The immediate effect of using building blocks is to move us farther away from our objective. There must be strong compensating advantages before we will restrict ourselves to any particular building block. The advantages of building blocks are:

  • Larger size. This means lower precision positional devices can satisfactorily manipulate the MBB.
  • More links between MBBs. As discussed in the next section, more linking groups on each building block implies more links between building blocks, greater stiffness (better bracing) and greater ease in forming three dimensional structures.
  • Greater tolerance of contaminants. Larger building blocks can have greater interfacial area, thus permitting the use of multiple weak bonds between building blocks (instead of fewer stronger bonds). As the particular pattern of interfacial weak bonds can be quite specific, two building blocks will bind strongly to each other while other molecules will bind weakly if at all. This principle, taken directly from self-assembly, is of great help in improving the ability of MBBs to tolerate dirt and other contaminants. This specificity also improves the ability of the MBBs to link even when positional accuracy is poor.
  • More accessible experimentally. While theoretical proposals clearly show the great potential of positional control when applied to very small building blocks (and even, under appropriate circumstances, individual atoms), the requirement for high precision and the intolerance of contaminants makes these proposals experimentally inaccessible with existing capabilities. MBBs can be relatively easy to synthesize and more tolerant of positional uncertainty and contaminants during assembly.
  • Ease of synthesis. Experimentally accessible MBBs must be synthesizable. As lower-strain structures are easier to synthesize, and polycyclic structures provide greater strength and stiffness, very low strain polycyclic structures (as in, e.g., diamond or graphite) are likely to be common in good MBBs. An exception to this general rule might be the construction of deliberately strained MBBs to facilitate the construction of curved surfaces (which would otherwise create strain in the inter-building-block linking groups) and to stabilize the cores of dislocations.
  • A larger design space. Perhaps the greatest advantage of MBBs is their vast number. As we increase their size the number of possible MBBs increases exponentially, giving us a combinatorially larger space of possibilities from which to select those few MBBs that best satisfy our requirements. While making it easier to satisfy our primary design constraints (ease of synthesis, number and specificity of inter-building-block linking groups, etc), this also makes it easier to satisfy secondary objectives such as non-flammability, non-toxicity, an existing literature, ability to work in multiple solvents as well as vacuum, tolerance of higher temperatures, etc.

In the next several sections, we discuss the characteristics and desirable properties of MBBs. In the section title “Proposals for MBBs” we consider some specific molecular structures that exemplify these properties. In the following section, we consider linking groups that can be used to connect some of the proposed MBBs. These include dipolar bonds, hydrogen bonds, transition metal complexes, and more traditional amide and ester linkages.

Following the discussion of MBBs and how to link them, we discuss higher-level strategies for making structures from them. The most obvious distinction is between subtractive synthesis (removing MBBs you don’t want from a larger crystal) and additive synthesis (adding MBBs you want to a smaller workpiece). The use of these two approaches places somewhat different requirements on the MBBs.

The goal of making larger MBBs might also be achieved by making them from smaller MBBs. The section on “starburst crystals” discusses an approach to this which might permit the synthesis of very large MBBs (perhaps ten nanometers or larger).

Finally, we consider what we might want to make from MBBs. If our objective is to implement positional assembly, then the most obvious thing to build is some sort of positional device. Other target structures, less ambitious than a complete positional device but which would be of use in a positional device, might be synthesized sooner as part of a longer term program.

Linking groups

MBBs can be characterized by the number of linking groups. More linking groups are generally better, as they more easily let us make stiff three dimensional structures. On the other hand, more linking groups tend to make the MBB harder to synthesize.

MBBs with three linking groups readily form planar structures because three points define a plane. Graphite, formed from sp2 carbon atoms which bond to three adjacent neighbors, is a planar structure that is quite strong and stiff in two dimensions but which, like paper, is readily folded through a third dimension. Just as paper can be formed into tubes to improve its stiffness, so can graphite be formed into tubes (often called bucky tubes).

MBBs with three linking groups could, like sp2 carbon, form planar structures with good in-plane strength and stiffness, but would be weak and compliant in the third dimension. While this problem could be reduced by forming tubular structures, stiff structures made using this approach would have to be made from many MBBs, as small numbers of MBBs (too few to form tubular structures) would lack stiffness.

MBBs with four linking groups not in a common plane are convenient for building three dimensional structures (much as the four bonds in a tetrahedral sp3 carbon atom allow it to form a stiff, polycyclic three-dimensional diamond lattice).

MBBs with three linking groups can be paired, each member of the pair sacrificing one linking group to form the pair. The pair of MBBs effectively has four linking groups (two available linking groups being provided by each member of the pair). Particularly if the four resulting linking groups are non-planar, the pair can be viewed as a single MBB with four linking groups. In this somewhat roundabout fashion, MBBs with three linking groups can form three dimensional structures much as MBBs with four linking groups.

MBBs with five linking groups can form three dimensional solids. For example, an MBB might have three in-plane linkage groups with inter-linkage-group angles of 120°; and have two out-of-plane linkage groups, both of which are normal to the plane (one linkage group pointing straight up, the other straight down). Such an MBB could form hexagonal sheets by using the three in-plane linkage groups, (each MBB corresponding to a single carbon atom in a sheet of graphite) but would also be able to link together adjacent sheets by using the two out-of-plane linking groups. The unit cell would have hexagonal symmetry.

MBBs with six linking groups can be connected together in a cubic structure, the six linking groups corresponding to the six sides of a cube or rhombohedron. MBBs with six linkage groups can naturally and easily form solid three dimensional structures in the same fashion that cubes or rhomboids can be stacked.

Buckyballs (C60) have now been functionalized with six functional groups (Hutchison et al., 1999; Quin and Rubin, 1999), opening up the possibility of using them as molecular building blocks for the construction of three dimensional structures.

An MBB with six in-plane linkage groups can form a particularly strong planar structure or sheet. The conformation of the sheet would depend only on the length of the inter-MBB links, and not on any ability of the MBB to maintain two linkage groups at some specific angle. As the distance between linked MBBs can often be controlled more effectively (the stretching stiffness of the link can be higher) than the angle between adjacent linkage groups (the bending stiffness is usually lower), this structure can be significantly stiffer in-plane than a planar structure formed from similar MBBs with three linkage groups.

Cubic or hexagonal close packed crystal structures are very stiff, involving 12 linking groups from each MBB. These structures can be described as follows: two very stiff sheets (six linkage groups in-plane) can be laid on top of each other. Each MBB in the upper sheet can be linked to three MBBs in the lower sheet (which form the vertices of a triangle). This arrangement can be repeated with a third, fourth, and more sheets. Six linkage groups connect each MBB to six in-plane neighbors, three linkage groups connect each MBB to three MBBs from the plane below, and three linkage groups connect each MBB to three MBBs from the plane above. The major advantage of this type of MBB is that the stiffness of the whole structure depends only on the stretching stiffness of the links between MBBs and not on the angular stiffness between adjacent linkage groups. This can be useful when the angular stiffness is poor, but the stretching stiffness is good.

MBBs with four linking groups can be paired, each member of the pair sacrificing one linking group to form the pair. The pair of MBBs effectively has six linking groups (three available linking groups being provided by each member of the pair). The pair can be viewed as a single MBB with six linking groups. This again leads naturally to unit cells that are cubic or rhomboid, but with each unit cell comprising two MBBs. This is similar to the primitive unit cell of diamond, which has two carbon atoms.

In summary, MBBs with two linking groups form three dimensional structures only with difficulty and only by using indirect and complex methods. MBBs with three linking groups readily form planar structures, which are strong and stiff in the plane but bend easily, like a sheet of paper, unless rolled into tubular structures to improve stiffness. They can also be used (although somewhat less naturally) to directly form three dimensional solids with a unit cell having four MBBs. MBBs with four linking groups quite naturally form strong, stiff three dimensional solids in which the unit cell is composed of two MBBs (as in diamond). MBBs with five linking groups can readily form strong, stiff three dimensional solids in which the unit cell is composed of six MBBs. MBBs with six linking groups readily form strong, stiff three dimensional solids in which the unit cell is composed of a single MBB. They can also form very stiff sheets if all linkage groups are in-plane, though this arrangement sacrifices stiffness out-of-plane. MBBs with twelve linking groups can form very strong and stiff three dimensional solids.

While MBBs can have any number of linkage groups, MBBs with fewer linkage groups are usually (though not always) more readily synthesized. If we seek an MBB with the least number of linkage groups that can still readily form strong, stiff three dimensional structures, then MBBs with four linkage groups are quite attractive. A high symmetry structure with four linkage groups will have tetrahedral symmetry (with an inter-linkage-group angle of approximately 109°). Much of the discussion in this paper is about specific tetrahedral MBBs.

Self assembled versus positionally assembled MBBs

The design criteria for self assembled MBBs differ in many fundamental respects from the design criteria for positionally assembled MBBs. For example, solubility constraints on positionally assembled MBBs are minimized. MBBs intended for self assembly in solution are usually soluble to permit them to explore differing orientations and positions with respect to each other, eventually settling on an energetically and entropically favored configuration. This solubility constraint is often non-trivial to satisfy and can greatly limit the range of MBBs that can be used.

In contrast, MBBs for positional assembly need not be soluble and do not even need a solvent: they can be used in vacuum. Like bricks, they can be picked up and moved to the desired location whether they are soluble or not.

If two MBBs can bond strongly to each other in two or more different configurations, then the self assembly process will randomly select from among these multiple configurations and produce a random clump of MBBs rather than any specific desired arrangement. For this reason, MBBs for self assembly often use multiple weak bonds, rather than a few strong bonds. Any particular weak bond can be broken by thermal noise. Only when the action of multiple weak bonds is combined does the resulting configuration of MBBs remain stable. Configurations that simultaneously enable multiple weak bonds are relatively rare, and so it is easier to design MBBs with multiple weak bonds that self assemble into a single desired structure.

While the use of strong bonds in self assembly is possible, positional assembly can more readily use MBBs that form a few very strong bonds. Inappropriate interactions between positionally assembled MBBs are prevented by the simple expedient of keeping them away from each other. When two MBBs are brought together, their orientations are controlled to prevent inappropriate bond formation. Thus, selective control over bond formation is achieved through positional control, rather than by designing the MBBs to be selective in bond formation. This approach permits the use of highly reactive MBBs that would be entirely inappropriate for self assembly.

The disadvantage of highly reactive MBBs is that they must be positionally controlled at all times. They cannot be allowed to mix randomly at any time, as this would cause them to rapidly form unusable clumps. While achievable, this imposes a number of constraints that are more difficult to meet with today’s systems.

An alternative approach is to use protecting groups that cover or alter the highly reactive linkage groups. These protecting groups would then be removed when two positionally assembled MBBs were joined. The use of protecting groups is common in chemical synthesis, though the concept of selectively removing a protecting group from a single molecule by using positional control is still novel. Selective photoactivation of molecules within a region comparable in size to the wavelength of light is well known and used commercially. From the perspective of nanotechnology, regions that are hundreds of nanometers in size are very large, making optical approaches rather imprecise when viewed from the perspective of the desired long term objectives.

Positionally assembled MBBs must be held, while self assembled MBBs need not be held. This implies that positionally assembled MBBs must have “handles” by which they can be gripped. While it is possible in some cases to use the linkage groups of the MBB as handles, these linkage groups might well be intended to irreversibly form strong bonds. Because it is essential in positional assembly both to hold the MBB and to let go, such an MBB must be able to form reversible attachments to the positional device. Ideally, this would be done using a variable affinity binding site which has two states: bound and unbound. The tip of the positional device first binds to the MBB. Then it positions the MBB with respect to some workpiece under construction, to which the MBB bonds. Finally, the positional device releases the MBB. Tweezers serve this function: when closed they can grasp an object (high affinity), when open they release the object (low affinity). While there are many other designs for variable affinity binding sites (Merkle, 1997b), tweezers are widely applicable and illustrate the basic concept.

Pragmatically, the greatest advantage of self assembled MBBs is the extensive literature on self assembly and the extensive set of existing experimental techniques that have been used to self assemble some impressively complex structures. Positional assembly at the molecular scale, by contrast, is still in its infancy. For example, the self assembly of DNA into complex molecular structures has made remarkable strides (Seeman, 1994), and has been used to make a truncated octahedron (Zhang, 1994). To quote Seeman and coworkers (Seeman et al., 1997):

There are several advantages to using DNA for nanotechnological constructions. First, the ability to get sticky ends to associate makes DNA the molecule whose intermolecular interactions are the most readily programmed and reliably predicted: Sophisticated docking experiments needed for other systems reduce in DNA to the simple rules that A pairs with T and G pairs with C. In addition to the specificity of interaction, the local structure of the complex at the interface is also known: Sticky ends associate to form B-DNA. A second advantage of DNA is the availability of arbitrary sequences, due to convenient solid support synthesis. The needs of the biotechnology industry have also led to straightforward chemistry to produce modifications, such as biotin groups, fluorescent labels, and linking functions. The recent advent of parallel synthesis is likely to increase the availability of DNA molecules for nanotechnological purposes. DNA-based computing is another area driving the demand for DNA synthetic capabilities. Third, DNA can be manipulated and modified by a large battery of enzymes, including DNA ligase, restriction endonucleases, kinases and exonucleases. In addition, double helical DNA is a stiff polymer in 1-3 turn lengths, it is a stable molecule, and it has an external code that can be read by proteins and nucleic acids. [references omitted from this quote]

The great drawback of self assembly, that it produces weak and compliant structures, can likely be adequately dealt with for transitional systems by careful design and post-modification of the self assembled structure to increase strength and stiffness. While the stiffness of DNA is good in comparison with most other polymers (Hagerman, 1988), it is still poor when compared with bucky tubes, graphite, diamond, silicon, and other “dry” nanotechnology materials.

Abstract properties of tetrahedral MBBs

Tetrahedral positionally assembled MBBs appear to be an attractive alternative, readily forming strong, stiff three dimensional structures while at the same time being simple enough that they can be synthesized. Before considering any specific tetrahedral MBB, we first consider some of their abstract properties.

First and foremost, the linkage groups will have particular properties. Of crucial concern are the conditions under which the links are made, and the extent to which inappropriate links are possible. If a specific functional group, call it R, bonds readily with other functional groups of type R (as is true for radicals), then the MBB cannot be kept in solution without rapidly forming undesired clumps. These limitations can be overcome by the use of protecting groups (or otherwise introducing some barrier to reaction) although this adds the additional requirement that the protecting groups be removed before an MBB is added to a growing workpiece.

A second type of linkage will involve two distinct functional groups, call them A and B. Functional groups of type A will readily bond with functional groups of type B, but A will not bond to A and B will not bond to B. The Diels-Alder reaction (Krummenacker, 1994) is a good illustration of this kind of functional group. The diene and dieneophile (corresponding to functional groups of type A and type B) will bond to each other, but not to themselves. They also bond to little else, and so can be used in most solvents (or in vacuum) and in the presence of impurities. As there are no leaving groups, the reaction itself does not introduce any possibly undesired contaminants.

A second advantage of the A-B functional groups is their increased tolerance of positional uncertainty. Consider two types of MBBs, type A and type B. Type A MBBs have four linkage groups of type A, while type B MBBs have four linkage groups of type B. Type As cannot link with other type As, nor can type Bs link with other type Bs. When type A and type Bs are combined in the diamond (or actuallyzinc-blende or wurtzite (hexagonal)) crystal lattice, they alternate. Each A is surrounded by four Bs, and each B is surrounded by four As.

In both the zinc-blende and wurtzite structures, there are no cycles of length five but many cycles of length six. That is, if we traverse a path from one MBB to another along the links between them, we will never find that we have completed a cycle and returned to the starting MBB without including at least six MBBs along the path. Clearly, a cycle with an odd length (such as five) would imply that either two As were linked or that two Bs were linked. This is forbidden by the nature of the A-B building blocks.

If, however, we were using MBBs of type R, which can readily link to each other, than a path of length five would be possible if the geometry of the MBBs was sufficiently distorted. Such a distortion might occur if the linking groups were insufficiently stiff, and permitted Rs near the edge of the crystal to come into contact with each other. This is exactly what happens on the diamond (100) surface, which forms strained dimers from adjacent carbon atoms which, if they were part of the bulk, would be separated by an additional carbon atom between them.

The use of A-B MBBs eliminates the possibility of odd cycles, and particularly cycles of length five. However, it does not eliminate cycles of length four, which could in principle occur if the geometry were sufficiently strained. As the strain required to achieve a cycle of length four is greater than the strain required to achieve a cycle of length five, R MBBs in a diamond lattice will more readily produce undesired cycles of length five than similar A-B MBBs will produce undesired cycles of length four.

This principle can be extended by introducing more types of functional groups, D, E, F, …. The more types of functional groups, the more strained the geometry must be before an incorrect link can be formed. In the limit, an arbitrary finite structure composed of a fixed number of MBBs arrayed in a regular lattice similar to diamond (or related structures) would have functional groups all of which were distinct and unique. Self assembly of such a structure would occur when the MBBs were mixed, and positional assembly would not be required at all. One method of providing a very large number of distinct types of functional groups is with DNA. There are many short DNA sequences that are selectively sticky, and which would bond only to the appropriate complementary sequence. Experimental work linking gold nanoparticles with DNA suggests this approach is experimentally accessible (Mirkin, 1996).

This illustrates a more general point: self assembly and positional assembly are endpoints on a continuum. As positional accuracy becomes poorer and poorer, “positional” assembly becomes more like self assembly. The techniques used in self assembly to ensure accurate assembly despite positional uncertainty can be gradually introduced into positional assembly as accuracy degrades.

This also illustrates the importance of stiffness in the linkage groups. The stiffer the linkage groups, the less likely that links will be formed where they shouldn’t be. In the limit, sufficiently stiff linkage groups would entirely prevent incorrect structures from forming. Provided the linkage groups have an appropriate orientation, the resulting structures will be unstrained (e.g., tetrahedral sp3 carbon atoms form unstrained bonds in diamond).

Proposals for MBBs

In this section, after reviewing the desirable properties of MBBs, we discuss some specific proposals for MBBs.

MBBs should be stiff, strong, and synthesizable with existing methods. Stiffness and strength are attributes derived from many strong bonds. Polycyclic molecules are usually stronger and stiffer than molecules without cycles (linear or tree structured molecules). Unstrained structures are usually easier to synthesize than strained structures. A good MBB is therefore likely to be polycyclic, with many strong, almost unstrained bonds. Given that bond-bond angles are often 120° (trigonal) or 109° (tetrahedral), we are likely to see hexagonal planar structures (as in graphite), or diamond and related structures. It should therefore come as no surprise that MBBs that resemble bits of graphite with appropriately functionalized edges, or bits of diamond with appropriately functionalized surfaces, are good candidates for MBBs. The diamond lattice in particular can be modified by substitution of carbon by elements from column IV: silicon, germanium, tin, or lead. Edge or surface atoms on the MBB can be chosen from columns III, V or VI, as appropriate (or alternatively the surface atoms can simply be hydrogenated).

Adamantane (hydrogens omitted for clarity)
1,3,5,7-tetraaza-adamantane (methenamine) (hydrogens omitted for clarity)

This line of reasoning leads fairly directly to molecules like adamantane: a stiff tetrahedral molecule which can incorporate heteroatoms and can be readily functionalized. Composed of 10 carbon atoms the Beilstein database (see www.beilstein.com) lists over 20,000 variants, supporting the idea that this family of molecular structures is large, contains many readily synthesized members, and has enough “design space” to provide solutions able to satisfy the multiple constraints imposed on a “good” molecular building block. This conclusion is further supported by (Fort, 1976), who surveyed adamantane chemistry.

A few molecules in this class which have been synthesized include adamantane; 1,3,5,7-tetrasila-adamantane; 1,3,5,7 tetrabora-adamantane (not yet synthesized, though 1-bora-adamantane has been synthesized); 1,3,5,7-tetraaza-adamantane (more commonly known as methenamine, readily synthesized and with a variety of commercial applications (Budavari et al., 1996)); 2,4,6,8,9,10-hexamethyl-2,4,6,8,9,10-hexabora-adamantane; and many others.

Tetramantane (hydrogens omitted for clarity)

Larger bits of diamond that have been synthesized include diamantane (C14H20), triamantane (C18H24), and even tetramantane (C22H28). Pentamantanes (C26H32) and hexamantanes (C30H36) occur naturally in some deep gas deposits (Schoell and Carlson, 1999; Dahl et al., 1999) but are not readily accessible in the laboratory.

Other small stiff structures that might be used as the basis for building blocks include cyclophanes, iceanes (small pieces of “hexagonal diamond,”), buckyballs, buckytubes, alpha helical proteins (Drexler, 1994), and a host of others.

An aside on “bond strength”

Bond “strengths” are typically measured in units of energy. Kcal/mol is common in the chemical literature, though electron volts, joules (more commonly atto joules (10-18 joules) or zepto joules (10-21joules)), Hartrees (atomic units often used in quantum chemistry software) and calories are all used as well. Conversion tables are commonly available. One convenient web page which lists some constants and conversion factors common in nanotechnology (and provides links to other sources) is at http://www.zyvex.com/nanotech/constants.html

In common (non-chemical) usage, “strength” refers to a maximum force, not an energy. Energy and strength are not the same, as the Newtonian equation relating work and force is Work = Force times Distance. Knowing the energy does not tell us the force that must be applied unless we also know the distance over which that force must work. In chemistry, a reasonable approximation to the stretching potential between two bonded atoms is the Morse potential:

U(x) = De[1 – e-b(x-r0)]2

Where U(x) is the potential energy of the system as a function of the separation x between the two bonded atoms, De is the “bond strength” as an energy, e is 2.71828…, r0 is the equilibrium distance (minimum energy distance), and b is a parameter which, along with De, determines the stiffness ks of the bond. As ks is readily determined from the vibrational frequency and the mass of the vibrating atoms, and De (with some adjustment for the zero point vibrational energy) is determined from chemical data about the bond strength, the parameter b can then be determined using the formula (Drexler, 1992):

b = sqrt(ks/ (2 De))

If we pull on the bond with a large enough steady force, it will eventually break. This occurs at a force of b De / 2. Using these equations, and knowing vibrational frequencies, atomic mass, and “bond strength” as an energy, we can compute the actual force required to break a bond. The force required to break a single carbon-carbon bond is ~6 nanonewtons. As the “strength” of the dipolar bond measured as an energy is almost one order of magnitude less, and the stiffness is not substantially less, we would expect that a force of roughly 1 nanonewton would be sufficient to break a dipolar bond.

The use of energies to measure bond strengths is appropriate if we expect that thermal noise is the disruptive force that will break bonds. The time until a bond is thermally disrupted is given by: tbreak = t0eDe/kbT

where kb is Boltzmann’s constant, T is the temperature in Kelvins, and t0 is a constant characteristic of the particular system (on the order of 10-13 for “typical” bonds). As can be seen, bonds with an energy “strength” significantly in excess of thermal noise will not be disrupted by thermal noise for a very long time.

Possible linking groups

As mentioned earlier, polymer chemistry has developed an enormous arsenal of functional groups that can link monomers together. The major drawback from the current perspective is that polymers made from monomers with only two linking groups tend to be floppy — rather than stiff, well defined three dimensional structures. (While proteins can fold into three dimensional structures, the process is indirect). By contrast, tetrahedral MBBs with four functional groups (to take one example) that can link to four other MBBs could be built into very stiff structures. Positional assembly of such MBBs would potentially enable the synthesis of an enormous range of structures. Thus, we wish to increase the number of linking groups per monomer.

How might we link adamantane-based MBBs? (hydrogens omitted for clarity)

How, then, might we link together two adamantane-based MBBs? One possibility which illustrates the concept is the use of dipolar bonds between nitrogen and boron. This is motivated less from any existing common polymer than from the observation that the simplest, stiffest and most direct method of linking two MBBs is to form a bond between two atoms, one atom from each adamantane. While radicals could be used, they suffer from certain drawbacks (clumping during synthesis, for example). The dipolar bond, on the other hand, permits synthesis of the B and N MBBs separately. While much stronger than hydrogen bonds, dipolar bonds are weaker than normal covalent bonds. Their strength can vary substantially, generally in a range from ten to a few tens of kcal/mol.

A central 1,3,5,7 tetrabora-adamantane MBB linked to four surrounding 1,3,5,7-tetraaza-adamantane MBBs.

If we use 1,3,5,7 tetrabora-adamantane and 1,3,5,7-tetraaza-adamantane as “B” and “N” building blocks, then each linking atom (N or B) is bonded to three other atoms in its MBB, thus providing a stiff support. The class of structures that can be formed includes the kind of structures typical of (for example) silicon carbide, where alternating silicon and carbon atoms are each bonded to four neighboring atoms of the other type. The B and N type adamantane-based MBBs are, in essence, larger versions of the same concept.

While 1,3,5,7 tetrabora-adamantane has not been synthesized, DFT calculations using a 6-311+G(2d,p) basis set show the molecule is a minima on the potential energy surface (Halls, 1999, private communication). Further, DFT calculations using a somewhat smaller 6-31G(d) basis set show that a dimer composed of an N and a B building block connected by a dipolar bond is also a minima on the potential energy surface with an enthalpy of formation of about 20 kcal/mol (ZPE corrected) (Halls, 1999). While boron with three single bonds is normally planar, it is strained by the tetrahedral nature of the adamantane cage. Stabilization of the boron atoms in the tetrahedral (rather than planar) bonding pattern by suitable electron donor groups (e.g., NH3) should increase the stability of the building-block plus four-donor-groups complex.

Hydrogen bonds

Hydrogen bonds are common in biological systems. They are relatively weak, on the order of 2-5 kcal/mol, but involve straightforward and widely practiced chemistry and can provide reasonable strength when several are combined (Watson et al., 1987). Two carboxcylic acids form a dimer via hydrogen bonds to each other with a ΔH of -14.1 to -16.4 kcal/mol in gas phase (Jones, 1952). If we use adamantane 1,3,5,7 tetracarboxylic acid (four COOH groups at the four trigonal carbon atoms of adamantane) as an MBB, each MBB can readily form eight hydrogen bonds to adjacent MBBs in the crystal if we assume that the MBBs are arranged like the carbon atoms in diamond. However, the resulting crystal structure would have large empty spaces. Experimental determination of the crystal structure (Ermer, 1988) shows five interpentetrating diamondoid lattices, thus effectively filling the large voids that a single diamondoid lattice would create.

Addendum added January 24, 2002: A theoretical possibility would be cyclohexane-1,3,5/2,4,6-hexacarboxylic acid (see figures at left, the graphic showing the energetically preferred all equatorial isomer). This MBB has six linking groups and each linking group could have two hydrogen bonds. While the all cis isomer — cyclohexane-1,2,3,4,5,6-hexacarboxylic acid with three axial and three equatorial groups — has been synthesized and is available commercially, it does not form an obvious crystal structure in which 12 well aligned hydrogen bonds can form. By contrast, while cyclohexane-1,3,5/2,4,6-hexacarboxylic acid has not been synthesized there is a theoretical crystal structure which forms 12 well aligned hydrogen bonds and has no large voids. Seven such MBB’s arranged in the appropriate structure are shown in the following figure:

Whether or not this theoretical crystal structure would actually form has not been experimentally determined. No obvious alternative structure would permit good alignment of all 12 hydrogen bonds. The pdb file for a cluster with eight MBB’s is here. There are many readily imaginable variants that have the same or a similar motif.

Addendum added March 15th 2002: “Acid B [all equatorial cyclohexane hexacarboxylic acid] is formed from acid A [all cis cyclohexane hexacarboxylic acid, commercially available] by heating with hydrochloric acid, …” (English translation of German patent by Badische Anilin and Soda-Fabrik Aktiengesellschaft, Convention Application No. 2212369, filed March 15 1972).

A paper that sheds tangential light on the possible crystal structure of “Acid B” is “The crystal structure of mellitic acid, (benzene hexacarboxylic acid)” by S.F. Darlow, Acta Cryst. (1961) 14, pages 159-166. This related molecule forms crystal with a structure as one might expect from the discussion here, differing largely in that it forms layers — a two-dimensional rather than a three-dimensional network of hydrogen bonds.

Tridentate complexes with transition metals

If the six edge atoms in adamantane are replaced with oxygen, then each “face” of the resulting tetrahedron will expose three oxygen atoms, each of which has one of its two lone pairs oriented towards that face. This opens up the possibility of a tridentate complex with an appropriate transition metal. A transition metal which could form a complex with six ligands (octahedral symmetry) could then form two tridentate complexes with the two faces from two neighboring building blocks. Substitutions in the frame of the adamantane cage could alter the spacing and type of the three donating groups (e.g., sulfur instead of oxygen) to permit the tuning of the building block for specific transition metals. This method would orient the other faces of the two building blocks appropriately for a diamond lattice (recall that the C-C-C-C torsion angle in the diamond lattice is n*120° + 60°, i.e., staggered rather than eclipsed; n is an integer).

Six linkage groups using adamantane

Adamantane has four atoms at the vertices of the tetrahedron, and six atoms along the edges of the tetrahedron. These six edge atoms could also be used to link the building blocks together. This would increase the number of links between building blocks (from four to six) thus strengthening the attachment of each building block to the whole. If we just think of carbon-carbon double bonds between adamantanes, the structure would be cubic with a unit cell consisting of 8 adamantane sub-groups. The larger number of linkage groups permitted by this approach might make weaker links more attractive. Hydrogen bonding (Watson, 1987) might prove effective, particularly if small clusters of OH groups could effectively be added despite the obvious steric problems.

Making larger building blocks from smaller ones

Larger building blocks are useful from at least two perspectives: they are easier to manipulate and their larger surface areas provide more sites to bind to other building blocks. While starburst dendrimers let us build large molecular structures from simple building blocks, the resulting structure is specified topologically but can be quite variable structurally.

Using two building blocks that alternate to form a crystal, a somewhat related but more structurally specific process (which we might call starburst crystals) would be to start with a single building block of type A and link it to as many building blocks of type B as possible (typically four or six). This might be done by adding a dilute solution of A to a concentrated solution of B, and then separating the ABnresults, where n corresponds to the number of linkage groups on A, under the assumption that all such linkage groups will be saturated with B building blocks. If the A building block is designated A0, then we might call the A building block surrounded by B building blocks A1.

This process can be repeated, the A1 building block can be mixed into a concentrated solution of A building blocks, adding an additional layer to the growing crystal and producing A2. A2 can be mixed into a concentrated solution of B building blocks, producing A3.

The critical difference between this process and the growth of a starburst dendrimer is that starburst crystalization adds building blocks only at those sites which extend the crystal structure. Thus, a new building block added in the Nth layer might bind to two or even three building blocks from layer N-1. Rather than exponential growth in the number of steps, this process has a growth rate that is cubic in the number of steps and the structure that results can be viewed as that part of a crystal that includes all crystal elements within a certain distance from some center.

If a building block added to layer N links to two building blocks from layer N-1, then it must link to the right two building blocks. If the links between building blocks are sufficiently stiff, this is not a problem. The new building block in layer N can only link appropriate pairs of building blocks from layer N-1. However, if the linkages are too floppy, the new building block might link to an incorrect pair of building blocks from layer N-1, producing an incorrect result. Preventing this requires either control over the geometry of growth (the bonds between building blocks cannot be too floppy) or selective control over linkage formation (different chemistries could be used for the formation of different links, even from the same building block; or linkage groups could be protected and de-protected).

Additive and subtractive synthesis

Given the building blocks and their natural tendency to form a zinc-blende crystal structure (the dipolar bond between two building blocks prefers the staggered rather than the eclipsed form — leading to a structure similar to diamond but with alternating building-block types), the set of structures that can be built include contiguous pieces of crystal with specific building blocks either present or absent. This state of affairs can be reached by one of two alternate routes: start with nothing and add building blocks until the desired structure is complete (additive synthesis), or start with a block which is too large and remove building blocks until the desired structure is reached (subtractive synthesis).

One method of additive synthesis is to add invididual building blocks (rather than groups of building blocks) one at a time. Using positional assembly, this requires a method of grasping and releasing the individual building blocks. As the building blocks already have four sites designed to bind to the complementary building block, these sites could be used to “grip” the building blocks while they were positioned. The tip of the positional device would need to be specifically designed to bind to the building blocks strongly enough to hold them while they were being positioned and oriented, but weakly enough that they could be released when the positional device was withdrawn from the workpiece under construction (or, alternatively, the tip could undergo some change to reduce its binding affinity for the building blocks).

Using substractive synthesis, undesired building blocks could be removed by scraping them away. The major advantage of this approach is that the tip of the positional tool need not bind to the building blocks, and therefore a much wider range of tip structures would be acceptable. Orientation requirements for the tip would also be relaxed. Force applied to a single building block on the surface would break it free from the workpiece. Provided that the bonds holding the building block together were significantly stronger than the bonds between building blocks, whole building blocks would be removed from the workpiece (rather than fragmenting the building blocks).

Subtractive synthesis has another advantage: adding building blocks one by one will on occasion produce situations where the new building block is bound to the workpiece by a single bond (the first building block to be added on a (111) surface, for example). Because it is held by only a single bond, the building block will not be as well bound to the rest of the structure and would have a higher probability of falling off before further building blocks could be added. This is of particular concern when weak bonds are being used between building blocks.

By contrast, substractive synthesis can leave intact all bonds that will be present in the final (desired) structure. If the final structure has been designed so that each building block is held in place by at least two bonds, then at every point during synthesis every building block that will be kept will be held in place by at least two (and often three) bonds. As the probability that a building block will break away from the workpiece is an exponential function of the depth of the potential energy well in which the building block finds itself, and as the depth of this well is doubled when two bonds hold it in place as compared with only a single bond, this difference can be significant.

One of the intriguing aspects of subtractive synthesis is the remarkably wide range of potential building blocks that could be used. Virtually any large, reasonably stiff and reasonably compact organic molecule that remains crystalline at reasonably high temperatures could be used. Adamantane itself, for example, melts at 268°C. (Weast, 1989). Such building blocks would be held together primarily by van der Waals forces, which would increase as the size of the building blocks increased. Precise modification by an SPM in UHV (Ultra High Vacuum) would seem feasible provided the tip was sharp in comparison with the size of the building block. The primary concern would be that building blocks on the surface (rather than in the bulk) might be so weakly bound that they would leave the surface. This could in general be dealt with by lowering the temperature, but a more careful search through the space of possibilities for building blocks that remained bound to the surface at room temperature might prove simpler. It might also be possible to find building blocks that could be selectively removed without the use of UHV, further simplifying the experimental procedure.

What to build?

Given a building block, what might we build? While our long term goals must be to build complex molecular machines, in the nearer term we will pursue the construction of key components. One possibility would be a set of molecularly precise tweezers (the use of carbon nanotubes as molecular tweezers has recently been experimentally demonstrated (Kim and Lieber, 1999)). Conceptually simple, a pair of molecularly precise tweezers could be picked up and manipulated by a larger pair of less perfect tweezers. The molecularly precise tweezers would provide well defined surfaces to interact with the part being manipulated.

A second obviously desirable structure would be a joint, which would provide one degree of rotational freedom and essentially no other degrees of freedom. The feasibility of sliding surface joints would depend in large part on the precise nature of the building blocks, but there is in general no problem in designing bearings with sliding surfaces (Merkle, 1993b). If the surfaces are not otherwise attractive to each other (e.g., hydrogen terminated carbon) then well designed bearings should have a small barrier to sliding motion. Bearings made from curved (strained) structures should be feasible at some scale, regardless of the building block, because some degree of strain is always tolerable.

Alternatively, multiple single-linkage-group bearings could be aligned. Two building blocks with intercalating layers could have single rotationally free linkage groups between facing layers. As the layers are molecularly precise, perfect alignment of such single-linkage-group bearings from successive layers would be feasible, thus permitting a molecular bearing of tolerable strength (at least at the molecular scale) to be built.

Finally, a more traditional door-hinge type of joint could be built by using intercalating layers from the two halves of the joint. Rather than attempting to strain the building blocks to provide smooth surfaces, relatively large holes (many building blocks in diameter) could be made which were aligned from layer to layer. A tubular pin (possibly made from strained building blocks, or possibly of some other type, such as buckytubes) could then be inserted through the holes, in the expecation that the smooth surface of the pin would be sufficient to support hinge rotation.

While the use of strained building blocks is feasible, it would also be possible to use building blocks that were “pre-strained.” For example, if a single edge atom in adamantane were changed from C to Si, the resulting building block would no longer be exactly tetrahedrally symmetric. Appropriately “malformed” building blocks could be used on specific crystal surfaces to relieve strain of a particular type. In addition, dislocations could be introduced into the structure. Special building blocks, designed specifically to relieve strain at the core of the specific dislocation, could be used to insure the stability (and feasibility) of the dislocation structure.

Rotary joints are of major importance for positional devices. It is possible to make a Stewart platform using nothing but appropriate rotary joints between otherwise rigid blocks. The design is left as an exercise for the reader — though we note here that each of the six struts in a Stewart platform must support two degrees of freedom at each end, much as a universal joint, and one degree of rotational freedom along the axis of the strut (Merkle, 1997c). Powering movement of the platform by moving the ends of the struts opposite the platform is a separate issue that is not dealt with by this design — though almost any powered one-degree-of-freedom movement of the “free” end of the strut would be sufficient.

Conclusions

The manufacture of molecular machines using positional assembly requires two things: positional devices to do the assembly, and parts to assemble. Molecular building blocks, made from tens to tens of thousands of atoms, provide a rich set of possibilities for parts. Preliminary investigation of this vast space of possibilities suggests that building blocks that have multiple links to other building blocks — at least three, and preferably four or more — will make it easier to positionally assemble strong, stiff three dimensional structures.

Adamantane, C10H16, is a tetrahedrally symmetric stiff hydrocarbon that provides obvious sites for either four, six or more functional groups. Over 20,000 variants of adamantane have been synthesized, providing a rich and well studied set of chemistries.

As positional assembly of molecules has only recently been recognized as a feasible activity, prior research in this area has been limited. No serious barriers to further progress have been identified, quite possibly because serious barriers do not exist. Progress will, however, require substantial further research