Basic Theory of Nuclear Magnetic Resonance Mark Betten, Calvin College Abstract— This paper describes the theory and applications of nuclear magnetic resonance (NMR) and how it applies specifically in scientific instruments. The papers discuses and explains the mathematical descriptions of the phenomenon, the electromagnetic theory explaining it, the subatomic particles that perform in the real world, and finally some of the major uses of NMR. Two specific examples are used as examples in understanding the fundamentals of NMR and the science of spectroscopy. Index Terms— I. INTRODUCTION: PRELIMINARY BACKGROUND A. What is NMR? Nuclear magnetic resonance (NMR) spectroscopy is the study of molecular structure through measurement of the interactions of an oscillating electromagnetic field with a collection of nuclei immersed in a strong external magnetic field [1]. The nuclei are the central parts of the atom which are assembled into molecules by bonds formed by electron orbital overlaps. An NMR spectrum can provide detailed information about molecular structure, both static and dynamic. Without NMR such acquisition of data would be extremely difficult, if not impossible. The phenomenal of the nuclei of certain atoms behaving strangely in the presence of strong external magnetic fields was first noticed by P. Zeeman [2]. In 1902 he was awarded the Nobel Prize in physics for this discovery [3]. However, it was not until 1952 that two physicists, Felix Bloch and Edward Purcell (Fig. 1.), were able to create a practical machine to observe the nuclear Zeeman Effect [4], [5]. Since the 1950’s the NMR spectroscopy has revolutionized the way chemistry, biochemistry, and biology is studied. NMR has become arguably the single most widely used technique for elucidation of molecular structure [1]. But to comprehend the NMR, a few fundamental principles of physics must be understood. B. Overview of Electromagnetic Radiation “All spectroscopic techniques involve the interaction of matter with electromagnetic reaction” and measuring the response of the matter to a given electromagnetic condition [1], [6]. The light rays that allow eyes to see constitute only a narrow range called the visible region of the electromagnetic radiation spectrum [7]. Each electromagnetic ray, as seen in Fig. 2, can be visualized as a composition of two orthogonal waves that oscillate exactly in phase with each other [1], [8]. This means that each wave reaches the peaks and nodes at the same points. One of the waves is the electric field vector (E) oscillating in one plane, and the other wave is the magnetic field vector (B) oscillating in the plane perpendicular to the electric field. Fig. 2. Electromagnetic wave with electric vector, E, and magnetic vector, B. Both fields exhibit uniform periodic motion. The axis along the abscissa can have units of length or time The electromagnetic waves consist of two independent parameters or properties, wavelength (λ) and the maximum amplitude (E0 and B0) [9]. However, the intensity of the wave is proportional to the square of its amplitude. Thus, given that the all electromagnetic radiation travels at a constant velocity c (3.00x1010 cm s-1 in vacuum), the wave can be described as having a frequency ν, which is the inverse of the peak-to-peak time t0: v Fig. 1. F. Bloch and E. Purcell (left to right) Manuscript for Calvin College Engr. 302 Course on May 19, 2004. Mark Betten is an engineering graduate from Calvin College with a concentration in Electrical and Computer. He is currently working for the VanAndel Research Institute in Grand Rapid, Michigan. (email: mbette27@calvin.ede) 1 t0 (1) where t0 is measured in seconds and v has units of cycles per second or hertz (Hz) [1], [9]. Realizing that the electromagnetic waves travel a distance of λ in t0 seconds, a second more important relationship can be deduced: c t0 v (2) The equation above shows that wavelength and frequency are not independent of each other, but are inversely proportional [9]. A wave of high frequency has a short 1 wavelength while a wave of low frequency must have a long wavelength. The electromagnetic spectrum (Table 1) ranges from the very small wavelength and high frequency of cosmic rays to the extremely large wavelengths and very low frequency of radio waves [1]. TABLE 1. Electromagnetic Spectrum and Properties In addition to its wave-like behavior, electromagnetic radiation also displays behavior characteristic of particles. A particle, or quantum, of radiation is called a photon. In essence, a photon can be considered a discrete packet of energy that is directly proportional to its frequency and can be summarized by the following relationship [9]: (3) E hv where h is Planck’s constant with a value of 6.63x10-34 Js per photon (or 3.99x10-13 kJsmol-1). Since the usual strength of a chemical bond is about 400 kJmol-1, electromagnetic wave energies above the visible region in Table 1 have more than enough energy to photodissociate (break) chemical bonds, while waves below the visible region cannot typically break molecular bonds [1]. The frequency that NMR spectroscopy uses is in the radio frequency region which also is the same frequency range that communication signals between radios and televisions [10]. Normally, the NMR uses frequencies from 200 to 759 MHz, which are at the low end of the energy scale (Table 1), but happens to precisely correspond to the amount of energy that is needed for molecular resonance and measurement [1]. C. Classical Interaction of Radiation with Matter Having briefly examined the properties of electromagnetic radiation, the parameters that control the interaction between waves and particles of matter can be examined. The three primary types of interaction of interest to spectroscopitists are absorption, emission, and scattering [10]. During an absorption interaction, the photon disappears and its entire energy is transferred to the particle that absorbed it. The resulting particle with excess energy is said to be in an excited state [9]. From its excited state, the particle can relax back to its ground state by emitting a photon (at a different wavelength than the incoming photon) which carries off the excess energy. Radiation is scattered when the direction of a traveling photon is shifted by some angle. The shift is due to the result of passing too close to a resonating molecule. If the frequency of the photon is unchanged the scattering is described as elastic. However, if the frequency has changed (inelastic scattering), this indicates that there must have been some exchange of energy between the photon and the particle [1]. In NMR spectroscopy, the main measurements are only absorption and emission of radio frequency radiation. Quantum mechanics, the field of physics that deals with energy at the atomic level and defines the rules that describe the probability for a photon to be absorbed or emitted under a given set of circumstances. But classical physics states a very important requirement shared by all forms of absorption and emission spectroscopy: for a particle to absorb (or emit) a photon, the particle itself must first be in some sort of uniform periodic motion with a characteristic fixed frequency. Most important, the frequency of the motion must exactly match the frequency of the absorbed (or emitted) photon: vmotion vphoton (4) Fig. 3. Energy Gap This above fact, which at first glance might appear to be an incredible chance, is actually quite reasonable. If a photon is to be absorbed, it energy, which is originally in the form of the oscillating electric and magnetic fields, must be transformed into energy of the particles’ motion. This transfer of energy can take place only if the oscillations of the electric and/or magnetic fields of the photon can constructively interfere with the oscillations of the particle’s electric and/or magnetic fields. If the energy hv of the photon is equal to the oscillation energy gap for the particle as seen in Fig. 3, only then can there be a transfer of energy to particle from the photon. When such a condition exists, the system is said to be in resonance, and only then can the act of absorption take place [1]. At this point one might thing that the frequency-matching requirement places a heavy constraint on the types of absorption processes that can occur. After all, how many kinds of periodic motion can a particle have? The answer is that even a small molecule is constantly undergoing many types of periodic motion. Each of its bonds is constantly vibrating and the whole molecule and some if its individual parts are rotating in all three dimensions. The electrons are circulating through their orbital. And each of these processes has its own characteristic frequency and it own selection of rules governing absorption [11]. All of the above forms of microscopic motion are what is called intrinsic, which means the motion takes place all by itself, without intervention by any external agent. However, it is possible under certain circumstances to induce particles to engage in additional forms of periodic motion. Still to achieve resonance, one needs to match the frequency of the induced motion with that of the incident radiation. 2 For example an ion (or any charged particle) follows a curved path as it moves through a magnetic field and if one was to carefully adjust the strength of the magnetic filed, the ion will follow a perfectly circular path, with characteristic fixed frequency that depends on it mass, charge, velocity, and strength of the magnetic filed (Fig. 4). Matching this characteristic cyclotron frequency with incident electromagnetic radiation of the same frequency ultimately can lead to absorption, and this is the basis of a technique known as ion cyclotron resonance (ICR) spectroscopy [1]. impossible to answer with complete precision [1], [11]. In 1927, W. Heisenberg, a pioneer of quantum mechanics stated his uncertainty principle: There will always be a limit to the precision with which we can simultaneously determine the energy and times scale of an event. Fig. 4. Charged particle moving in perpendicular magnetic field. D. Uncertainty and Time Scale. To take a photograph of moving object, one needs to know the shutter speed of the camera that must be set to avoid blurring the image. Therefore, the faster the object is moving, the shorter the exposure times to capture the motion of the object. Similar considerations in spectroscopy must be also taken. Suppose one was looking at a particular dot which could turn white and black every 1 s, as illustrated in Fig. 5. If one was to take a picture of this dot with a shutter speed of 10 s, the photograph of the dot would appear to be gray, because it would capture the average of the two colors of the dot. But if on decreased the exposure time to perhaps 0.01 s, the photograph would show a black one or white one dot. Thus to capture the individual colors, black or white in this case, the exposure time must be significantly shorter than the cycle time of the color change. In a similar fashion, frequency, analogous to the shutter speed, must be high enough so that the motion of the particle can be captured. In addition, electrical hardware and computational power must be fast enough to capture the image of the particle. Fig. 6. W. Heisenberg Stated mathematically, the product of the uncertainties of the energy and time can never be less than h . (5) Et h Thus, if we know the energy of a given photon to a high order of precision, we would b unable to measure precisely how long it takes for the photon to be absorbed. However, there is a useful generalization one can make by substitution of h, giving: (6) t 1/ v ,where v is the uncertainty in frequency. As a result, the time required for a photon to be absorbed ( t ) must be approximately as long as it takes one cycle of the wave to pass the particle [1], [11]. The length of time t0 in Fig. 2, is nothing more than 1 / v . This result is not surprising if one considers that the particle would have to wait through at least one cycle before it could “sense” what the radiation frequency was. This allows an idea of the order-ofmagnitude of how fast the “shutter” (frequency) speed of the NMR must be in order to “freeze” a given molecular transition. II. MAGNETIC PROPERITES OF NUCLEI Fig. 5. Freezing a transient condition requires a “fast” shutter. There are many types of molecular changes, that is, molecules constantly undergo some sort of reversible reorganization of their structures. If absorption of the photon is fast enough, one should be able to detect both “black” and “white” forms of the molecules. But if the absorption process is slower than the interconversion, the NMR will only detect some sort of the time-averaged structure, making the structure harder to decipher. The situation therefore boils down to the question: How long does it take for particle to absorb a photon? Unfortunately, such a question is A. The Structure of an Atom The compounds examined by NMR are composed of molecules, which are themselves aggregates of atoms. Each atom (Fig. 7) has some number of negatively charged electrons whizzing around a tiny dense bit of positively charged matter called the nucleus [12]. The size of an atom is the volume of space that the electron cloud occupies. However, >99.9% of the mass of an atom is concentrated in its nucleus, though the nucleus occupies only one trillionth the atom’s volume. Even the nucleus can be further dissected into other fundamental particles, including protons and neutrons, not to mention a host of other sub-nuclear particles that help hold the nucleus together and give nuclear physicists something to wonder about. 3 Fig. 8. The shape of the orbital based on the quantum number. Fig. 7. Bohr model of an atom. 1) Composition of the Nucleus The number of protons in an atoms nucleus (Z, the atomic number) determines both the atom’s identity and the charge on its nucleus. In the periodic table of elements, the atomic number of each element is shown to the right of it chemical symbol. Every nucleus with just one proton is a hydrogen nucleus while every nucleus with six protons is a carbon nucleus, and so on. Yet, careful examination a large sample of hydrogen atoms, finds that not all their nuclei are identical. It is true that all have just one proton, but they differ in the number of neutrons. Most hydrogen atoms in nature (99.985%, to be exact) have no neutrons (N=0), but a small fraction, 0.015%, have one neutron (N=1) in addition to the proton. These two forms are the naturally occurring stable isotopes of hydrogen and they are given the symbols 1 H and 2H, respectively. The leading superscript is the mass number (A) of the isotope, which is the integer sum of Z and N: (7) AZ N 2 The isotope H is usually referred to as deuterium (D), or heavy hydrogen, but most isotopes of other elements are identified simply by their mass number. The atomic mass listed for each element in the periodic table is a weighted means, the fractional abundance of each isotopes times its exact mass, summed over all naturally occurring isotopes [1], [9], [11]. 2) Electron Spin Before going further into the properties of the nucleus, a close examination of the electrons is needed to understand electron spin. Just like protons, electrons also exhibit wave and particle like properties. Each electron wave in an atom is characterized by four quantum numbers. The first three of these numbers (n, l, and m) can be taken as the electrons’ address and describes the energy (n), shape (l), and orientation (m) of the volume the electron occupies in the atom. All three numbers define a specific volume of the electron which is called an orbital [9]. Fig. 8 and 9 show the various electron orbitals as defined by the specific quantum numbers [13]. Table 2 shows the allowable quantum numbers for each energy level [1]. Fig. 9. The orientation of the orbital based on the quantum number. The fourth quantum number is the electron spin quantum number s, which can assume only two values +½ or –½. The Pauli exclusion principles states that the no two electrons in an atom can have exactly the same set of four quantum numbers. In other words, if two electrons occupy the same orbital, they must have different spin quantum numbers, one +½ and the other –½ or visa versa. Therefore, no orbital can possess more than two electrons and the only if their spins are paired (having opposite values). Table 2. Summary of Allowable Quantum Numbers Because the electrons can be regarded as a particle spinning on an axis, it has a property called spin angular momentum. Further, because the electron is a charged particle (Z=-1), the spinning give rise to a magnetic moment (μ) represented by the vector in Fig. 10. Such particles are described as having a magnetic dipole [6]. The two possible values of s correspond to the two possible orientations of the magnetic moment vector in an external magnetic field “up” (in the same direction as the external field) or “down” (in the opposite direction to the external field). The two spin states 4 are degenerate (i.e., having the same energy) in the absence of an external magnetic field. Moreover, if all the electrons in an atom are paired (i.e. each orbital contains two electrons), all up spins are cancelled by the down spins, so the atom as a whole has zero magnetic moment [1], [9]. Fig. 10. Two possible orientations of the magnetic moment μ of a spinning electron in an external magnetic field Bo. When unpaired electrons are immersed in an external magnetic field (such as Bo), the two states (s=+½ or s=–½) are no longer degenerate. An electron with its magnetic moment oriented opposite the external field has a lower energy than an electron with its magnetic moment oriented with the external field (Fig. 10). It is the inter-conversion of the two spin state that is centrally important to the technique known as electron paramagnetic resonance spectroscopy [1]. 3) Nuclear Spin The proton is a spinning charged (Z=1) particle as well and so it too exhibits a magnetic moment. And as with the electron, its magnetic moment has only twp possible orientations that are degenerate in the absence of an external magnetic field. To differentiate nuclear spin states from the electronic spin states, the proton spin quantum number m is used. Thus for a proton, m can assume values of only +½ or -½ and can be described as having a nuclear spin I, (I= +½ or -½). Since nuclear charge is opposite in sign that of the electron, a nucleus whose magnetic moment is aligned with the magnetic field (m= +½) has the lower energy (Fig. 11). Fig. 11. Two possible orientations of the magnetic moment μ of a spinning proton in an external magnetic field Bo. Although neutrons are uncharged subatomic particles, neutrons also display magnetic moments and show a spin of I= ½ [10]. As previously noted, Zeeman found only certain isotopes give rise to multiple nuclear spin state when immersed in an external magnetic field. This is because only isotopes with an odd number of protons (odd Z) and/or and odd number of neutrons (odd N) posses nonzero nuclear spin. Nuclei with zero nuclear spin (when Z and N are even) have zero nuclear magnetic moment and cannot be detected by NMR methods [1], [3]. Thus on of the limitations to NMR is that only certain isotopes can be detected. What follows from the above observation is the importance of parity. The reason that parity (odd or even number) of protons and neutrons is so important is that a proton spin can only pair (cancel) another proton spin, but not a neutron spin, and vice versa. This rule organizes every elemental isotope into one of three groups [1]. However, remember that different isotopes of the same element can have different nuclear spins, some of which are detectable by NMR and others which are not. a) Group 1: Nuclei with Both Z and N Even In these atoms, the nuclei has all proton spins paired and all neutron spins paired, resulting at a nuclear spin of zero (I=0). Such nuclei are invisible to the NMR. Some examples of these atoms are 12C, 16O, 18O, and 32S [1]. b) Group 2: Nuclei with Both Z and N Odd In these atoms, the nuclei have odd number of unpaired proton (I=½) spins and odd number of neutron (I=½) spins, so that the net magnetic spin must be nonzero integer (i.e., the integer must be a multiple of 2½). Atoms with these nuclei are detectable by the NMR. Examples include 2H (I=1), 10B (I=3), 14N (I=1), and 50V (I=6) [1], [9]. c) Group 3: Nuclei with Even Z and Odd N These nuclei must have an even number of proton spins (all paired) and an odd number of unpaired neutron spins, or vice versa. This causes the net magnetic spin to be an odd integer multiple of ½, which allows the nuclei to be detected by the NMR. Some examples include include 1H (I=½), 12B (I= 1½), 15N (I=½ ), 17O (I=2½), 19F (I=½), 29S (I=½) and 31P (I=½) [1], [10]. B. The Nucleus in the Magnetic Field 1) The Nuclear Zeeman Effect As stated before a nucleus with a nuclear spin I adopts 2I+1 nondegenerate spin orientations in a magnetic field. The states separate in energy, with the largest positive m value corresponding to the lowest energy (most stable) state. It is this separation of states in the magnetic field that is the essence of the nuclear Zeeman Effect. The energy of the ith spin state (Ei) is directly proportional to the value of mi and the magnetic field strength Bo (that is the energy is quantized in units of hBo / 2 ), Ei mihBo 2 (8) In this equation h (Planck’s constant) and π are the usual 5 constants seen in physics. γ is called the magnetogyric ration which is a proportionality constant characteristic of the specific isotope being examined. The minus sign in the equation follows from the convention of making a positive m correspond to lower (negative) energy. Fig. 12 and Fig. 13 graphically shows the variation of spin state energy as a function of the magnetic field strength for two different nuclei, one with I=½ and I=1. Fig. 12. Nuclear Zeeman effect. A nucleus with I=½. uniform periodic motion prior to excitation. Luckily, quantum mechanics require that the magnetic moments are actually not statically aligned exactly parallel or anti-parallel to the external magnetic field as Fig. 10 and 11 suggest. Instead, the nuclei are forced to remain at a certain angle Bo which causes them to “wobble” around an axis of the filed at a fixed frequency [9]. The reason for the “wobble” is analogous to the situation of a spinning top. If one observes a spinning top, it has spin angular momentum that prevents it form falling over and also causes it to wobble in addition to spinning. The periodic wobbling motion a top assumes in a gravitational field is called precession. In the same way the earth precesses on its axis, the nucleus does a similar process. In an exactly analogous frequency called the Larmor frequency (ω), which is a function solely of γ and Bo: (12) Bo The angular Larmor frequency, in units of radians per second, can be transformed into linear frequency ν by division by 2π, Bo (13) precession 2 2 The processional motion causes the tip of the magnetic moment vectors (either up or down) to trace out a circular path, as shown in Fig. 14. Also note that the procession frequency is independent of m, so that all spin orientations of a given nucleus precess at the same frequency in a fixed magnetic field. Fig. 13. Nuclear Zeeman effect. A nucleus with I=1. Notice that as the field strength increases, the difference in energy (ΔE) between any two spin states also increases proportionally. For a nucleus with I=½, the difference is E E ( m 1 / 2) E ( m 1 / 2) (9) hBo (10) [( 1 ) ( 1 )] 2 2 2 hBo (11) 2 Now the reason +½ or –½ were chosen for m and s seems clear. It is so that the change in energy can be in constant incremental terms. The magnetogyric ratio γ describes how much the spin state energies of a given nucleus vary with changes in the external magnetic field. Each isotope with nonzero nuclear spin has it own unique value of γ, but the magnitudes of γ depends on the units of the external field Bo [1], [6], [13]. 2) The Precession and the Lamor Frequency When nuclei are immersed in a magnetic field, the particle will adopt 2I+1 spin orientations, each with a different energy. But before the nuclei can absorb photons, it is assumed that they must have been oscillating in some Fig. 14. Precession of the magnetic moment in states I= ½. III. MATHEMATICS BEHIND NMR A. Fourier Transforms A Fourier transform is an operation which converts functions from time to frequency domains. An inverse Fourier transform (IFT) converts from the frequency domain to the time domain [14]. The FT is defined by the integral: (14) Think of f(ω) as the overlap of f(t) with a wave of frequency ω, (15) 6 This is easy to picture by looking at the real part of f(ω) only. important parts of the spectrometer are not that complex to understand briefly; and it is extremely helpful when using the spectrometer to have some understanding of how it works. Figure 16 shows the typical NMR machine. (16) The actual FT will make use of an input consisting of a REAL and an IMAGINARY part. Mx is considered the REAL input, and My as the IMAGINARY input. The resultant output of the FT will therefore have a REAL and an IMAGINARY component as well. In FT NMR spectroscopy, the real output of the FT is taken as the frequency domain spectrum. To see an esthetically pleasing (absorption) frequency domain spectrum, we want to input a cosine function into the real part and a sine function into the imaginary parts of the FT. This is what happens if the cosine part is input as the imaginary and the sine as the real This conversion is made using a mathematical process known as Fourier transformation. This process takes the time domain function (FID) and converts it into the frequency domain function (spectrum) as seen in Fig. 15. Fig.16. Typical NMR machine. Fig. 15. Fourier Transformation B. Convolutions To the magnetic resonance scientist, the most important theorem concerning Fourier transforms is the convolution theorem. The convolution theorem says that the FT of a convolution of two functions is proportional to the products of the individual Fourier transforms, and vice versa. If f( ) = FT( f(t) ) and h( ) = FT( h(t) ) then f( ) g( ) = FT( g(t) f( ) f(t) ) and g( ) = FT( g(t) f(t) )i IV. NMR HARDWARE NMR spectrometers have now become very complex instruments capable of performing an almost limitless number of sophisticated experiments. However, the really Broken down to its simplest form, the spectrometers consists of the following components [11]: An intense, homogeneous and stable magnetic field A “probe” which enables the coils used to be exited and detect the signal to be placed close to the sample. A high-power RF transmitter capable of delivering short pulses (oscillating magnetic field) A sensitive receiver to amplify the NRR signals A digitizer to convert the NMR signals into a form which can be stored in computer memory A “pulse programmer” to produce precisely timed pulses and delays A computer to control everything and to process the data. The following sections provide information how each part works. A. Static Magnetic Field Modern NMR spectrometers use persistent superconducting magnets to generate the static magnetic fields. Basically such magnet consists of a coil of wire through which a current passes, thereby generating a magnetic field. The wire is of special construction such that at low temperatures (usually less than 6K), the resistance goes to zero so that the wire is superconducting. Once the current is set running in the coil it will persist forever and generate a magnetic field without the need for further electrical power. Superconducting magnets tend to be very stable and so are very useful for NMR [11]. To maintain the wire in it superconducting state, the 7 coil is immersed in a bath of liquid helium. Surrounding this is usually a “heat shield” kept at 77K by contact with a bath of liquid nitrogen. This reduces the amount of expensive liquid helium needed, which often boils off due to the heat flowing tin from the surroundings. The whole assembly is constructed in a vacuum flask so as to further reduce the heat flow. The cost of maintaining the magnetic field is basically the cost of the liquid helium and liquid nitrogen. B. The Shims The lines in NMR spectra are very narrow – line widths of 1 Hz or less are not uncommon – so the magnetic field has to be very homogeneous, meaning that it must not vary very much over space. The reason for this is easily demonstrated by an example. Consider a proton spectrum recorded at 500 MHz, which corresponds to a magnetic field of 11.75 T. Recall that the Larmor frequency is given by o where 1 B 0 eq. x 2 is the gyromagnetic ratio (2.67x10^8 rad/sT for protons). We need to limit the variation in the magnetic field across the sample so that the corresponding variation in the Larmor frequency is much less that the width of the line, say by a factor of 10. With this condition, the maximum acceptable change in Larmor frequency is 0.1 Hz and so using the equation above, we can compute the change in the magnetic field as, the bore of the magnet. The small coil used to both excite and detect the NMR signal is held in the top of this assembly in such a way that the sample can come down from the top of the magnet and drop into the coil. Various other pieces of electronics are contained in the probe, along with some arrangements for heating or cooling the sample. The key part of the probe is the small coil used to excite and detect the magnetization. To optimize the sensitivity this coil needs to be as close as possible to the sample, but of course the coil needs to be made in such a way that the sample tube can drop down from the top of the magnet into the coil. Extraordinary effort has been put into the optimization of the design of this coil. The coil forms part of a tuned circuit consisting of the coil and a capacitor. The inductance of the coil and the capacitance of the capacitor are set such that the tuned circuit they form is resonant at the Larmor frequency. That the coil forms part of a tuned circuit is very important as it greatly increases the detectable current in the coil. Spectroscopists talk about “tuning the probe” which means adjusting the capacitor until the tuned circuit is resonant at the Larmor frequency. Usually we also need to “match the probe” which involves further adjustments designed to maximize the power transfer between the probe and the transmitter and receiver. Figure 11. shows a typical arrangement. 2 0.1 / 2.4 x10 9 T Expressed as a fraction of the main magnetic field this about 2x10-10. We can see that we need to have an extremely homogeneous magnetic field for work at this resolution. On its own, no superconducting magnet can produce such a homogeneous field. What we have to do is to surround the sample with a set of shim coils, each of which produces a tiny magnetic field with a particular spatial profile. The current through each of these coils is adjusted until the magnetic field has the required homogeneity, something we can easily assess by recording the spectrum of a sample which has a sharp line. Essentially how this works is that the magnetic fields produced by the shims are canceling out the small residual in homogeneities in the main magnetic field. Modern spectrometers might have up to 40 different shim coils, so adjusting them is a very complex task. However, once set on installation it is usually only necessary on a day to day basis to alter a few of the shims which generate the simplest field profiles. The shims are labeled according to the field profiles they generate. So, for example, there are usually shims labeled x, y and z, which generate magnetic fields varying the corresponding directions. There are more shims whose labels you will recognize as corresponding to the names of the hydrogen atomic orbital. This is no coincidence; the magnetic field profiles that the shims coils create are in fact the spherical harmonic functions, which are the angular parts of the atomic orbital. [11] C. The Probe The probe is a cylindrical metal tube which is inserted into Fig. 5. Typical Probe The two adjustments tend to interact rather, so tuning the probe can be a tricky business. To aid us, the instrument manufacturers provide various indicators and displays so that the tuning and matching can be optimized. We expect the tuning of the probe to be particularly sensitive to changing solvent or to changing the concentration of ions in the solvent. D. The Transmitter (Oscillating Magnetic Field) The radiofrequency transmitter is the part of the spectrometer which generates the pulses. We start with an RF source which produces a stable frequency which can be set precisely. The reason why we need to be able to set the frequency is that we might want to move the transmitter to different parts of the spectrum, for example if we are doing experiments involving selective excitation. Usually a frequency synthesizer is used as the RF source. Such a device has all the desirable properties 8 outlined above and is also readily controlled by a computer interface. It is also relatively easy to phase shift the output from such a synthesizer, which is something we will need to do in order to create phase shifted pulses. As we only need the RF to be applied for a short time, the output of the synthesizer has to be “gated” so as to create a pulse of RF energy. Such a gate will be under computer control so that the length and timing of the pulse can be controlled. The RF source is usually at a low level (a few mW) and so needs to be boosted considerably before it will provide a useful oscillating field when applied to the probe. The complete arrangement is illustrated in Figure 12. E. Digitizing the Signal A device known as an analogue to digital converter or ADC is used to convert the NMR signal from a voltage to a binary number which can be stored in computer memory. The ADC samples the signal at regular intervals, resulting in a representation of the FID as data points [11]. The output from the ADC is just a number, and the largest number that the ADC can output is set by the number of binary “bits” that the ADC uses. For example with only three bits the output of the ADC could take just 8 values: the binary numbers 000, 001, 010, 011, 100, 101, 110 and 111. The smallest number is 0 and the largest number is decimal 7 (the total number of possibilities is 8, which is 2 raised to the power of the number of bits). Such an ADC would be described as a “3 bit ADC”.ii The waveform which the ADC is digitizing is varying continuously, but output of the 3 bit ADC only has 8 levels so what it has to do is simply output the level which is closest to the current input level; this is illustrated in Figure 13. Fig. 6. The analog signal (a) is digitized in (b). Therefore, the numbers that the ADC outputs are an approximation to the actual waveforms. The approximation can be improved by increasing the number of bit. This gives more output levels. At present, ADC with between 16 and 32 bits are commonly in use in NMR spectrometers. The move to higher numbers of bits is limited by technical considerations. The main consequence of the approximation process which the ADC uses is the generation of a forest of small sidebands – called digitization sidebands– around the base of the peaks in the spectrum. Usually these are not a problem as they are likely to be swamped by noise. However, if the spectrum contains a very strong peak the sidebands from it can swamp a nearby weak peak Fig. 7. Typical Transmitter Configuration RF amplifiers are readily available which will boost this small signal to a power of 100 W or more. Clearly, the more power that is applied to the probe the more intense oscillating field will become and so the shorter the 90 degree pulse length. However, there is a limit to the amount of power which can be applied because of the high voltages which are generated in the probe. When the RF power is applied to the tuned circuit of which the coil is part, high voltages are generated across the tuning capacitor. Eventually, the voltage will reach a point where it is sufficient to ionize the air, thus generating a discharge or arc (like a lightening bolt). Not only does this probe arcing have the potential to destroy the coil and capacitor, but it also results in unpredictable and erratic oscillating fields. Usually the manufacturer states the power level which is “safe” for a particular probe [5]. F. The Receiver The NMR signal emanating from the probe is very small but for modern electronics there is no problem in amplifying this signal to a level where it can be digitized. These amplifiers need to be designed so that they introduce a minimum of extra noise (they should be low-noise amplifiers). The first of these amplifiers, called the pre-amplifier or pre-amp is usually placed as close to the probe as possible (you will often see it resting by the foot of the magnet). This is so that the weak signal is boosted before being sent down a cable to the spectrometer console. One additional problem which needs to be solved comes about because the coil in the probe is used for both exciting the spins and detecting the signal. This means that at one moment hundreds of Watts of RF power are being applied and the next we are trying to detect a signal of a few µV. We need to ensure that the high-power pulse does not end up in the sensitive receiver, thereby destroying it [7]. This separation of the receiver and transmitter is achieved by a gadget known as a diplexer. There are various different ways of constructing such a device, but at the simplest level it is just a fast acting switch. When the pulse is 9 on the high power RF is routed to the probe and the receiver is protected by disconnecting it or shorting it to ground. When the pulse is off the receiver is connected to the probe and the transmitter is disconnected.iii Some diplexers are passive in the sense that they require no external power to achieve the required switching. Other designs use fast electronic switches (rather like the gate in the transmitter) and these are under the command of the pulse programmer so that the receiver or transmitter are connected to the probe at the right times.iv V. NMR APPLICATIONS IN SCIENTIFIC RESEARCH A. Carbon 13 NMR Many of the molecules studied by NMR contain carbon. Unfortunately, the carbon-12 nucleus does not have a nuclear spin, but the carbon-13 (C-13) nucleus does due to the presence of an unpaired neutron. Carbon-13 nuclei make up approximately one percent of the carbon nuclei on earth. Therefore, carbon 13 NMR spectroscopy will be less sensitive (have a poorer signal to noise ratio) than hydrogen NMR spectroscopy [11]. that carbons that do not have directly bonded protons (i.e. carbonyls and quaternaries) have much longer relaxation times than protonated carbons [1]. REFERENCES [1] Macomber, Roger S. A Complete Introduction to Modern NMR Spectroscopy. New York: John Wiley & Sons, Inc., 1998. [2] http://en.wikipedia.org/wiki/Zeeman_effect [3] http://www.thebakken.org/library/books/20z.htm [4] http://www.aip.org/history/esva/catalog/esva/Purcell_Mi lls.html [5] http://www.aip.org/history/esva/catalog/esva/Bloch_Feli x.html [6] Rahman, Atta-ur. Nuclear Magnetic Resonance Basic Principles. New York: Springer-Verlag, 1986, pp 1, 7. [7] http://www.ntia.doc.gov/osmhome/allochrt.pdf [8] http://en.wikipedia.org/wiki/Electromagnetic_radiation [9] Chemistry Text Book [10] Hornack,, Joseph P. http://www.cis.rit.edu/htbooks/nmr/inside.htm [11] Keeler, James. Understanding NMR Spectroscopy. Cambridge: University Cambridge Press, 2002. [12] http://www.chemistry.mcmaster.ca/esam/Chapter_3/sect ion_3.html [13] http://hyperphysics.phy-astr.gsu.edu/hbase /quantum/vecmod.html [14] Lathi, B. P. Signal Processing and Linear Systems Figure 5.1. Typical C13 NMR. The sample was C5H7O2N When measuring carbon spectra, the main concern is usually signal to noise. You would expect higher field spectrometers to have a decisive advantage - for example a 500 Mhz spectrometer when compared to a 300 MHz spectrometer should have an advantage of (5/3) squared, or 2.8 times the signal to noise. However there are other considerations, including for example the type of probe. An indirect detection probe has the proton observe coil on the inside (that is, closer to the sample than the coil used for carbon). This improves the proton signal to noise, however if you use an indirect detection probe for directly observing carbon, the signal to noise will of course be worse than a standard probe which has the carbon coil on the inside. Regardless of the probe design, carbon and protons use different coils, and since the electronic circuit for the two nuclei is different it makes no sense to compare proton signal to noise on two instruments and extrapolate the results to carbon [2]. Also, signal to noise tests are usually performed by collecting a single scan on a concentrated sample, however this does not give the best indication of the results obtainable on "real" samples where the sample is scanned for several hours. When a sample is repeatedly pulsed, the relaxation times of the various carbons must be taken into consideration. Nuclei take longer to relax at higher fields, so the gain in signal to noise is less than expected. Also note 10