JEOL USA Blog

What Is Surface Analysis?

What Is Surface Analysis?

Surface analysis is an analytical technique to elucidate elemental composition, chemical state, and micro structure from material surface layer (several nm to several µm). As phenomena such as corrosion, wear, adhesion, and reactions that impact performance and reliability occur primarily on the surface, surface analysis is vital for material evaluation, quality control, and failure analysis.
For analysis, it is necessary to select the most suitable method, according to the sample state (target point, size, material, etc.) and analysis purpose.
What information we want to know? What is the material of the sample?
What is the range of information that we want to know? How deep?
Is it water-soluble? Does it react with solvents? Pre-treatment needed?
It is important to select the analytical method suitable for the purpose.

Types and features of surface analysis instruments

The figure below shows the comparison of typical surface analysis methods from various points of view, such as excitation source, detection signal, quantitativeness, whether the chemical state can be analyzed or not, sensitivity, handling of an insulator, and analysis capability of depth direction. It is important to understand the feature of each method and properly select the analysis method according to the purpose.
Analytical methods EPMA (WDS)/SXES/EDS AES XPS XRF SIMS
Excitation source Electron beam Electron beam X-ray X-ray Ion
Signal Characteristic X-ray Auger electron Photoelectron Fluorescence X-ray Secondary ion
Detectable element Be ~ (WDS, EDS)
Li (SXES, Windowless EDS)
Li ~ Li ~ C ~ H~
Quantitative analysis ×
Chemical state × Organic compound
Detection depth Several µm Several nm Several nm Several mm Several nm
Sensitivity Several ten ppm
(Mass concentration)
Several thousand ppm
(Atomic concentration)
Several thousand ppm
(Atomic concentration)
Several ten ppm
(Mass concentration)
Several ppm
(Atomic concentration)
Insulator ○ (Conductive coating)
Depth analysis ×
Strength Qualitative analysis
Quantitative analysis
Wide area ~ micro area analysis
Micro area analysis
Chemical bonding state analysis
Depth profile analysis
Insulator analysis
Chemical bonding state analysis
Depth profile analysis
Qualitative analysis
Thin film analysis
Trace element analysis
Organic substance analysis
Trace element analysis
Challenge Chemical bonding state analysis
(Strong at SXES)
Organic substance analysis
Wide area analysis
Insulator analysis
Organic substance analysis
Micro area analysis
Trace element analysis
Micro area analysis Qualitative analysis
Quantitative analysis
In this column, we explain surface analysis instruments that JEOL offers, such as XPS (photoelectron spectrometer), AES (Auger microprobe) , XRF (X-ray Fluorescence Spectrometer) , EPMA (Electron Probe Microanalyzer) with *standard wavelength-dispersive X-ray spectrometer, SEM+EDS (Scanning Electron Microscope+Energy Dispersive X-ray Spectrometer), and SXES (soft X-ray emission spectrometer) that can be installed to EPMA (WDS) and SEM.
We clearly explain each mechanism, its strengths and weaknesses in analysis, and key points for selecting the instrument.

Difference of analysis area/depth according to surface analysis instrument

XRF enables elemental analysis in the deepest and widest region. It is suitable for understanding the average composition of the entire bulk material and is utilized in qualitative/quantitative analysis in a wide field of view.
On the other hand, SEM + EDS and EPMA (WDS) can investigate the local elemental distribution by detecting x-rays that are generated in a local area of about several micrometers. SEM+EDS enables simultaneous evaluation of morphology and elemental analysis, while EPMA provides superior capabilities in more precise quantitative analysis and area analysis.
Moreover, AES and XPS makes it possible to obtain signals from the very shallow surface layer of about several nanometers deep. They are optimal for evaluating chemical state of the surface layer such as surface processing, contamination, and oxidization state.
Thus, to investigate the extreme surface at the nanometer scale, AES or XPS is suitable. For local analysis at the micrometer level, SEM combined with EDS or EPMA is appropriate. If the target is a wide area on the millimeter scale, XRF is the best choice. The appropriate instrument varies depending on the required analysis depth and the field of view size.

Difference in principles and detection signals of surface analysis instruments

As shown below, each instrument has a different excitation source (incident probe) and detection signal, and the information obtainable is different according to their features.

Points for selecting surface analysis method

In surface analysis, it is important to select the appropriate technique based on the properties of the specimen and the purpose of the analysis. For specimens that are susceptible to vacuum, such as biological or liquid specimens, methods like XRF, which can be performed under atmospheric pressure, or SEM equipped with a low-vacuum mode are effective. If the specimen can withstand a vacuum environment, more sensitive and higher-resolution techniques such as XPS, AES, or EPMA can also be considered. This section introduces the optimal analytical method for each purpose, along with the figures.

Qualitative/Quantitative Analysis

Area Analysis

State Analysis

Summary

Surface analysis is a technique to obtain key information that directs to performance and reliability of the instrument. This article explains the tips of instrument selection through the features of typical analysis methods and points of selection, strength by instrument, and concrete application examples.
JEOL Ltd. has product line-ups that can satisfy a wide range of needs from beginners to researchers, and that can be utilized with support from introduction to operation.

COSY/TOCSY Analysis│Interpreting spin correlations using 2D NMR

COSY/TOCSY Analysis│Interpreting spin correlations using 2D NMR

In this article, we explain the basic principles and analysis methods of COSY and TOCSY, which are representative techniques of 2D NMR. Starting from confirming correlations between adjacent protons using COSY, we introduce spin network analysis with TOCSY and analysis of carbohydrate using 1D and 2D TOCSY, with concrete examples.

What is COSY?

COSY(COrrelation SpectroscopY)is a basic method for visualizing couplings between adjacent 1Hs by using 2D NMR. Traditionally, homodecoupling of 1Hs was used to locate neighboring 1H atoms one by one. However, with the appearance of COSY, it became possible to analyze 1H-1H coupling correlations simultaneously over a broad range, dramatically improving the efficiency of structural analysis. Currently, COSY is widely used as an introductory 2D NMR method for confirming 1H-1H correlations.

COSY shows correlations between adjacent 1Hs - in other words, 1Hs that are 3 bonds apart. The spin coupling of 3 bonds apart is expressed as 3JHH. If the adjacent 1H is known, the connection of 1Hs in a molecule can be understood.

Fig.1 COSY spectrum of cinnamic acid cis-3-hexenyl ester
Fig.1 COSY spectrum of cinnamic acid cis-3-hexenyl ester

Fig.1 shows a COSY spectrum for the region corresponding to 1Hs at 1 to 6 in "cinnamic acid cis-3-hexenyl ester". As indicated above, when the correlation signals and 1H spectrum of 1D on both axes are connected, 5 correlations can be observed: 1-2, 2-3, 3-4, 4-5, and 5-6. The result allows us to infer that 1Hs from 1 to 6 have a neighboring structure each. In addition, the COSY allows us to understand the connection between carbon atoms indirectly from the coupling information of neighboring 1Hs. The information is extremely important for determining the substructure.

Fig.1 COSY spectrum of cinnamic acid cis-3-hexenyl ester
Fig. 2 Complex COSY spectrum
However, what happens when the COSY spectrum is as complex as the one in Fig. 2? Looking at the region highlighted in green, you can see that the signals overlap, making it hard to identify correlations. In this situation, moving forward with structure determination seems almost impossible.

For compounds like polysaccharides or cyclic molecules, where 1H signals densely appear in a narrow range as in Figure 2, relying on COSY alone can lead to a dead end in structural analysis.

One solution for the situation when the COSY does not adequately analyze is TOCSY. We will explain TOCSY in the next section.

What is TOCSY?

TOCSY(TOtal Correlation SpectroscopY)is a 2D NMR technique that allows you to visualize all 1Hs belonging to the same spin network at once. Because spin couplings propagate through the network, TOCSY reveals not only adjacent positions but also correlations across the entire spin system, making it highly effective for analyzing complex organic molecules, sugars, and amino acids. TOCSY is also known as HOHAHA(HOmonuclear HArtmann-HAhn spectroscopy), which essentially refers to the same experiment.

Fig. 3 Network of continuous 1H spin couplings

For example, suppose that there are atomic connections as shown in Figure 3. HA・HB・HC・HD are connected through spin couplings between neighboring 1Hs, and this connection is called spin system (spin network). The spin system does not propagate through quaternary carbons without attached H or 1H coupled via oxygen, so the network is interrupted at these points. TOCSY allows us to observe correlations between nuclei within the same spin system even if they are not directly coupled--for instance, between 1HA and 1HD. In TOCSY, the magnetization is transferred step by step: from HA to HB, then from HB to HC, and finally from HC to HD. In other words, magnetization moves through the spin system, connecting everything along the way. This magnetization transfer is called "relay". TOCSY observes the correlation signals based on this relay.

Fig. 4 Pulse sequences of COSY and TOCSY
Fig. 4 shows the pulse sequences for COSY and TOCSY. COSY uses two 90° pulses, while with the TOCSY, the second pulse is a spin lock pulse. The duration of this spin lock pulse is called "mixing time," which is a key parameter with TOCSY. By increasing the mixing time, you can observe correlation signals that have been relayed over longer distances within the spin system.

Mixing time and correlation signal of TOCSY

Fig. 5 Schematic diagram of TOCSY spectrum

Fig. 5 shows a schematic diagram of a TOCSY spectrum of a compound having two independent spin networks. Each spin network is expressed in yellow and in green. As shown, TOCSY observes correlation of all 1Hs belonging to each spin network. At that time, if signals like "A" and "a" within the primary spectrum on both axes appear away from the region where signals appear densely, the signals are easily classified separately in the networks of yellow and green, enabling confirmation of each spin network.

Fig.6 Schematic diagram of TOCSY (Correlation of A)

Next, let's look at how the correlation signal of A changes when the mixing time is varied. Figure 6 is an enlarged view of the correlation of A from the schematic TOCSY diagram in Figure 5. With a short mixing time, the magnetization of 1HA moves only to the neighboring 1HB, and the correlation signal B with 1HB appears. When the mixing time is further increased, the magnetization moves to 1HC, and correlation signal C appears. When the mixing time is increased even more, the magnetization moves to 1HD, and correlation signal D appeared. As a result, it becomes clear that there is a connection of 1HA - 1HB - 1HC - 1HD.

The longer the mixing time, the farther the magnetization can move. Therefore, if a sufficiently long mixing time is used, the magnetization can be transferred through all 1Hs within the same spin system, allowing the detection of correlation signals for all 1Hs in that spin system. Furthermore, by comparing spectra obtained with different mixing times, the sequential order of the 1H connections can also be known.

Example of analysis for sucrose using 2D TOCSY

20 mg / 0.6 mL D2O solution (400 MHz)

Here, we present an example of analysis for sucrose using 2D TOCSY. Sucrose is a disaccharide composed of glucose and fructose linked by a glycosidic bond. Sucrose contains a total of 22 1Hs, but since the sample is dissolved in heavy water, the 1Hs of the hydroxyl groups (-OH) are not observed due to deuterium exchange. Therefore, in this case, 14 1Hs are observed, excluding the 8 1Hs of hydroxyl group.

Fig.7 COSY spectrum of sucrose
Let us show you the COSY spectrum of sucrose in Fig. 7. The part highlighted in the red frame is enlarged and shown in Fig. 8.
Fig. 8 COSY spectrum of sucrose
In Figure 7, six correlation signals can be found relatively easily. However, it is difficult to interpret the correlations of signals appearing in areas where chemical shifts are close to each other, such as those indicated by the red circles in Figure 8. Many people may get stuck at this point in the analysis. Therefore, we will try using the TOCSY method.
TOCSY spectrum with mixing time of 20ms
TOCSY spectrum with mixing time of 150ms
Fig. 9 TOCSY spectra measured with mixing times set to 20 ms and 150 ms

Figure 9 shows the TOCSY spectra measured with mixing times set to 20 ms and 150 ms, respectively. First, we focus on the correlation signal of the 1H at the anomeric position of glucose, which appears at a chemical shift well away from other signals. As shown earlier in Figure 6, we will examine how the correlation signals change with different mixing times, to confirm the relay information from the 1H at the anomeric position. It can also be observed that more correlation signals appear as the mixing time increases. Furthermore, for regions where signals overlap, it is easier to compare by using sliced data rather than the 2D spectrum. Therefore, we will compare each 1D spectra obtained by extracting slices along the X-axis (in the area indicated by the green frame) for the confirmation of correlation signal of the 1H.

Figure 10. One-dimensional spectra sliced along the X-axis for the correlation signal of the 1H at the anomeric position of sucrose

The top spectrum in Figure 10 shows a conventional 1H spectrum. Below it are sliced data obtained with mixing times varied from 20 ms to 200 ms. The signal on the left corresponds to the Glu H-1, which we consider as the starting point of the relay. At a mixing time of 20 ms, only the correlation with the neighboring Glu H-2 appears, revealing only the connection between positions 1 and 2. As the mixing time increases, correlation signals appear sequentially, and eventually, the connections from position 1 to position 6 in the glucose moiety can be confirmed. While it was difficult to clearly identify the correlation partner of the Glu H-5 using COSY, TOCSY provided this information.

Next, let us examine the relay starting from the Fru H-3′. As with the sucrose moiety, Figure 11 shows the TOCSY spectra obtained with different mixing times.
TOCSY spectrum measured with mixing time of 20ms
TOCSY spectrum measured with mixing time of 150ms
Figure 11. TOCSY spectra measured with mixing times set to 20 ms and 150 ms
As before, we compare sliced data along the X-axis for a signal that appears in a region without significant overlap--this time focusing on the Fru H-3′.
Fig. 12 One-dimensional spectra sliced along the X-axis for the correlation signal of the H-3′ proton of fructose

The top spectrum in Figure 12 shows a conventional 1H spectrum, and below it are sliced data obtained with different mixing times. As before, we take the Fru H-3′ as the starting point of the relay. As the mixing time increases, correlations with the Fru H-4′, Fru H-5′, and Fru H-6′ appear, allowing us to confirm the connections within the fructose moiety.

In this way, TOCSY enables the separation of spin networks within a compound that contains multiple spin networks. By varying the mixing time, it is also possible to assess the distance from the starting 1H. Even when the COSY spectrum becomes complicated, using TOCSY provides significant assistance in determining the molecular structure.

Example of analysis for glucose using 1D TOCSY

1D TOCSY is the one-dimensional version of 2D TOCSY, and the basic principle is the same. While 2D TOCSY is used to analyze the spin network of the entire molecule, 1D TOCSY is more effective when you want to examine only the spin system to which a specific 1H belongs. 1D TOCSY observes the magnetization transfer--i.e., the relay--from a selectively excited 1H. By gradually increasing the mixing time, the propagation of the magnetization relay along the spin network can be tracked. For selective excitation, it is best to choose a 1H whose chemical shift is sufficiently separated from other signals. Another advantage of 1D measurement is that, compared to 2D measurement, it offers higher digital resolution and makes it easier to avoid signal overlap.

Now, let us present an example of glucose analysis.

Fig 13. 1H spectrum of glucose

Glucose exists in aqueous solution as α- and β-anomers. Therefore, the 1H NMR spectrum is a mixed spectrum of α-glucose and β-glucose, as shown in Figure 13. For selective excitation of a 1H signal, it is typically to choose 1H appearing at a lower field than others (in the case of sugars, 1H at the anomeric position). The result of a 1D TOCSY experiment by selectively exciting the anomeric 1H at position 1 of α-glucose is shown in Figure 14.

Vivamus sagittis lacus vel augue laoreet rutrum faucibus dolor auctor. Duis mollis, est non commodo luctus.
Fig.14 1D TOCSY spectrum of glucose

By varying the mixing time from 20 ms to 200 ms, it was possible to trace the spin network starting from the anomeric 1H at position 1 of α-glucose as the relay point. The bottom spectrum with a mixing time of 200 ms represents the extraction of only the spin network of α-glucose from the mixed spectrum of α-glucose and β-glucose. Furthermore, as explained earlier, 1D TOCSY provides better digital resolution than 2D sliced data. Thus, 1D TOCSY is more effective for extracting and confirming spin systems connected through spin couplings from crowded spectra. It is recommended to use 1D TOCSY and 2D TOCSY appropriately depending on the purpose.

Basics of NOE/NOESY: Causes and Solutions When NOE Is Not Detected

Basics of NOE/NOESY: Causes and Solutions When NOE Is Not Detected

NOE (Nuclear Overhauser Effect) is an important NMR measurement technique that can reveal spatial proximity relationships within molecules. In this column, we discuss the difference between difference NOE, 1D NOESY, and 2D NOESY, and how to select each method, the relaxation time that is key for analysis. The causes and solutions for cases when NOE is not detected are also covered. Moreover, we answer the frequently asked questions and explain the key points for effective use of NOE measurement.

What is NOE measurement?

What is NOE measurement?
NOE measurement (Nuclear Overhauser Effect) is a method to observe nuclei in spatially close distance. The NOE measurement is utilized to obtain stereochemistry of molecules, and distinguish isomers and confirm location of substituent group, for application of small molecular compound.

Principle of NOE

The NOE is a phenomenon where spatially close nuclei magnetically affect each other and cause an NMR signal intensity change.

1. Relation between nuclear spin and magnetic field

When placed under the influence of external magnetic field, the energy state of nuclear spin of proton splits (Zeeman splitting).

2. Saturation by radio wave

When radio wave continues to irradiate to a specific proton, the spin state of the proton "saturates". In saturation state, energy absorption and release are balanced, and the signal disappears.

3. Dipole - dipole interaction

When saturated proton magnetically interacts with another spatially close proton, the signal intensity of this proton changes. This interaction is classified as "bipolar-bipolar interaction".
Principle of NOE
Considering NOE, we think about two protons in one molecule as shown in the above figure.
When these protons are close to each other, the NOE is observed due to dipole interaction.
On the other hand, if the distance between protons is far, the NOE is not observed as there is no dipole interaction.
It is known that the dipole interaction is closely related to the inter-nuclear distance and molecular mobility, and if the distance between two nuclei is shorter than 6 Å, NOEs may be observed.
On the other hand, since the magnitude of the interaction is proportional to the inter-nuclear distance (r) to the minus sixth power (r-6), When the inter-nuclear distance increases, the interaction rapidly decreases, making it difficult to observe NOEs.

How to observe NOE

Different from the splitting by spin-coupling, the NOE is not observed with the NMR spectrum (1D, 2D NMR) that we have introduced so far.
Therefore, the following steps are required for the NOE measurement.
  • Irradiating radio wave to specific proton (excitation/saturation)
  • Observing signal intensity change of other protons
  • Judging if this change is coming from the NOE

Types of NOE Measurement

Types of NOE Measurement
There are three main types of NOE measurements.
  • Difference NOE: observe a steady state NOE by taking the difference in spectra with and without radio wave irradiation.
  • 2D NOESY: observe transient NOE by using a 2D spectrum.
  • 1D NOESY: selectively excite protons and observe transient NOE in one dimension.
Steady state NOE is to irradiate a radio wave for a long time, make a specific proton saturated, and observe the signal intensity change of other protons once the saturation state becomes steady.
Transient NOEs are observed during the mixing time after the spin state is temporarily changed by a short period of radio wave irradiation. Since the mechanisms by which NOEs are generated are different, the measurement conditions for each method must be determined by considering the characteristics of each.

Difference NOE Measurement

Difference NOE measurement is a method of checking the change in signal intensity of nearby protons by taking the difference between the spectra of a specific proton when irradiated with radio waves and when not irradiated with radio waves.
Difference NOE Measurement

As shown above, two protons, HI and HS are going to be considered. HI and HS are in spatially close distance and having a dipole interaction. (B) is 1H spectrum when HS was not irradiated.

Here, when a radio wave is irradiated to HS to saturate HS, the signal of HS becomes invisible. At this time, the signal intensity of HI becomes large due to the dipole interaction between HI and HS. This is the "Change due to NOE". Since the change of signal intensity by NOE is very small, the difference between (A) and (B) is calculated to make the difference clear. The NOE observed by difference NOE is called "steady state NOE". In general, for a small molecular compounds having high speed molecular motion in solution, when the signal change of irradiated proton is expressed downward, the NOE signal is observed as an upward signal. This is called "Positive NOE".

Actual example of difference NOE measurement is introduced, by using ethyl crotonate.
Difference NOE Measurement
We would like to determine whether the conformation of a double bond is cis or trans by using the difference NOE measurement. The proton of methyl group expressed as A is irradiated, and the related NOE is observed.

In the upper part of the above figure, the 1H spectrum of ethyl crotonate is shown. In the lower part, difference NOE spectrum when a proton of methyl group at A is saturated, is shown. This data is measured by using 400 MHz instrument and the NOE signals are observed upward at B and C.

Based on the obtained results, irradiation of the proton of the methyl group at A led to observable NOEs at B and C, indicating that ethyl crotonate has a trans conformation.

1D NOESY

1D NOESY is a method to selectively excite a specific proton and observe the transient NOE between the specific proton and the other spatially nearby protons as a one-dimensional spectrum. In this measurement, soft pulse (narrow bandwidth pulse) and pulsed field gradient are used to excite only the proton of interest, and the change in NOE over the mixing time is recorded. Since there is no need to subtract spectra as in difference NOE, spectra that are easy to analyze are obtained. 1D NOESY is suitable when you have a specific proton of interest or when you want to check the NOE in a simple way.
1D NOESY
Here, we present an example of a 1D NOESY measurement using ethyl crotonate. The (B) in the above figure shows 1D NOESY spectrum obtained by selective excitation of proton A, while (C) shows the spectrum obtained by selective excitation of proton C. Unlike difference NOE spectra, 1D NOESY does not show the residual signals caused by subtraction of two spectra obtained before and after radio wave irradiation, making the spectrum easier to analyze.
The interpretation method is similar to that of difference NOE: we check how other signals appear when the selectively excited proton signal is shown downward.
In example of ethyl crotonate, when proton A is selectively excited, positive NOEs are observed at B and C. Conversely, when proton C is excited, a positive NOE is observed only at A. These results also support the estimation that ethyl crotonate has a trans type stereochemistry. Thus, when determining stereochemistry with ambiguous spatial arrangements, it is important to measure 1D NOESY of the selective excitation of each part.

2D NOESY

2D NOESY is a method to observe interaction (NOE) between spatially close protons in a molecule as a two-dimensional spectrum. This experiment observes the appearance of how NOE changes (transient NOE) during the mixing time after temporarily changing the proton spin state. The obtained spectrum expressed the self-signals on a diagonal and correlated signals between spatially close protons. Since the 2D NOESY allows for comprehensive NOE correlation in the entire molecule, it is suitable for analysis of complicated structures and confirmation of entire signal assignment.
2D NOESY
From here, we will explain how to read 2D NOESY spectra. As shown in the above figure, when A and C, B and D are in close proximity to each other, such correlation signals appear. The data sliced at positions A and B (in blue dotted line) in X-axis direction, are equivalent to 1D NOESY spectra by selective excitation of A and B, respectively.
This figure shows 2D NOESY of ethyl crotonate. Looking at each sliced data, the signals at the selective excited part (in red) appear downward, while signals with NOE correlation (in black) appear upward (positive NOE).
When looking at the correlation signals in black, the following NOE correlations can be found:
A:B and C
B:A
C:A
D:E
E:D
In addition, focusing on the NOE correlations among A, B, and C, the results support the presence of the trans type conformation, consistent with the previously mentioned difference NOE and 1D NOESY results.

Easy guideline in selecting NOE measurement method

As we have explained, for an NOE measurement, it is possible to choose the method according to the purpose. We recommend that you begin with 1D NOESY which allows for easy condition setting and clear spectra to be obtained.
Target proton has been specified: 1D measurement
Want to see NOE across the molecule, or signal assignment is uncertain: 2D NOESY

Important factor for NOE measurement: Relaxation time

One of the most frustrating problems in NOE measurements is, without a doubt, "when no NOE signal is detected!" In such a case, please measure the longitudinal relaxation time (T1) of the target sample, first. Then, for difference NOE measurement, please set the irradiation time of the excitation pulse to be more than 5 times that of T1 of the target sample. Also, please set the mixing time with 1D/2D-NOESY, to be about 0.8 times that of T1 of the target sample.

Now, why do we need to consider the relaxation time of the target sample, for an NOE measurement? This is because NOE has a deep relationship with relaxation time, as NOE is a relaxation phenomenon based on the dipole interaction.
Here, we would like to provide a simple explanation about relaxation in NMR.
The above is the schematic diagram of energy levels of nuclear spin of spin number= 1/2 when it is placed in a static magnetic field. Nuclear spin takes up two energy states after undergoing Zeeman splitting by being affected by magnetic field. Stable state with low energy is called α state, while the one with high energy is called β state. When a radio wave having a frequency equivalent to this energy gap is irradiated, the spin in the α state absorbs the energy of radio wave and excited into the β state. When the spin number of the α state and the β state becomes the same by irradiation of a radio wave, energy absorption stops, which is called the saturation state. In the saturation state, the spin in the β state emits energy and returns to the α state, finally, it even returns to the thermal balanced state. The whole process is called relaxation. Then, a series of processes from excitation to relaxation, are called the NMR phenomenon that we are observing.
Two neighboring nuclear spins
The interaction works between two neighboring nuclear spins, which is the driving force for the relaxation explained earlier. In solution NMR, the dipole interaction is one of the major factors in promoting relaxation. Therefore, setting the parameter of relaxation time which has deep relationship with dipole interaction is an important key for successful NOE observation.

Cause and solution for undetectable NOE signals

The likely cause and action when NOE is not detected, are explained in the following 4 points.
  1. Optimize sample adjustment/measurement conditions
  2. Change molecular motion
  3. Change resonance frequency
  4. Change measurement method
First, check if 1. is optimized, and if the NOE is not observed, try 2 to 4.

1. Optimize sample adjustment/measurement conditions

Since the NOE signal is very weak, the result is greatly affected by the sample concentration and purity, and measurement condition setting.
  • Concentration: if the concentration is too high, relaxation tends to occur between molecules, weakening the NOE.
  • Caution against contamination of paramagnetic substances: dissolved oxygen and metal ions (Fe, Cu, etc.) promote relaxation, thus preventing NOE.
  • Setting of irradiation/mixing times: Based on T1(longitudinal relaxation time), irradiation time of more than 5 times, mixing time of about 0.8 times are guidelines.

2. Change molecular motion

The intensity of NOE also depends on molecular motion. The following graph shows the relationship between NOE intensity and molecular motion.
This curve was theoretically brought from the relational expression of NOE intensity and dipole interaction.

The vertical axis expresses the NOE intensity, while the horizontal axis expresses the product of observation frequency ω multiplied by correlation time of the molecule tC. tC is the time required for a molecule to make a full rotation in solution and is a parameter expressing the motion of the molecule. Which means, the smaller the molecule is, the shorter the tC becomes. The larger the molecule is, the longer the tC becomes.

The left shows the region of low molecule where tC is short and the motion of molecule is fast. Here, a positive NOE is observed. On the other hand, the right shows the region of polymer where tC is long and the motion of the molecule is slow. Here, the negative NOE is observed. Which means, the graph indicates that NOE intensity greatly changes depending on the motion of the molecule.

What matters in performing measurement is that there is a region where NOE becomes zero. When the molecule is close to the region where the product of multiplication of observation frequency ω and correlation time of the molecule tC is 1, the NOE becomes very weak or the NOE is not observed.

Therefore, if NOE is not observable even if measurement conditions are reviewed with your instrument, it is possible that the product of multiplication of observation frequency ω and correlation time of the molecule tC is in this region.

In such a case, changing ω or tC is a possible choice. However, changing ω means changing the measurement magnetic field, and it may be difficult to try it at once if you hold only one instrument. Therefore, we suggest that you consider changing tC , which is, changing the factor that can affect tC(measurement temperature and viscosity of a solvent). We recommend that you change the measurement temperature, first.

3. Change resonance frequency

Next, I will explain the case where the observation frequency is changed, which is, where the measurement magnetic field is changed.

The above figure shows the magnetic field dependence of NOE intensity and correlation time tC .The vertical axis represents NOE intensity, while the horizontal axis represents correlation time tC. Each curve corresponds to a magnetic field from 90 MHz to 920 MHz. The curves shift to the left as the measurement magnetic field increases. If we focus on the tC at which the NOE becomes zero, we can see that the point at which the signal intensity becomes zero shifts to the region of low molecules (where the molecules are in fast motion) as the magnetic field is increased. This means that under the same measurement conditions, the signal intensity of positive NOE becomes weaker at higher magnetic fields. For example, it can be said that a molecule with zero NOE in a 600 MHz instrument may have a positive NOE observed in a 300 MHz instrument. In addition, for the case of a molecule with about 1000 molecular weight, the possibility is high that it can be in the region in a circle where NOEs are extremely small, so attention is needed.

4. Change the measurement method

If NOE is not observed after trying methods 1 through 3 introduced so far, a measurement method called ROESY is used. ROESY is a NOE measurement in a rotational coordinate system.
The above figure is the 1D NOESY and 1D ROESY spectra of gramicidin S measured at 90°C. At this temperature, the NOE of gramicidin S becomes nearly zero, but using ROESY allows the observation of correlation signals. The NOE observed in ROESY is referred to as ROE. The greatest advantage of ROESY is that, as illustrated in the schematic diagram, ROE is always positive and does not have a zero region. You might wonder, "Then why not just use ROESY from the beginning?" However, ROESY requires more careful parameter settings and tends to produce unwanted signals. Therefore, ROESY is better used when NOESY does not work well.

Frequently Asked Question about NOE Measurement

In difference NOE measurements, if the irradiation position is not set at the peak maximum, spin coupling effects can occur, causing changes in the signal shape. Therefore, the irradiation position should be set precisely at the peak maximum.
In difference NOE measurements, when multiplets more than doublets are selectively irradiated and if the irradiation position is shifted from the peak maximum, the signal shapes are changed. The two spectra above show difference NOE results when the methyl group protons of ethyl crotonate are irradiated. Positive NOE signals are observed at B and C. If the irradiation position is shifted to around the center ppm of signal A instead of the peak maximum, a large, asymmetric, residual-like signal appears, as shown in the dotted area of the top spectrum. This is due to spin coupling interactions with proton B. Always select the irradiation position at the peak maximum in difference NOE measurements.
This phenomenon is known as indirect NOE (three-spin effect), and in small molecules, it can be observed when protons are arranged nearly linearly.
Normally, small molecules undergoing rapid motion in solution exhibit positive NOE signals. However, signals may sometimes appear in the same direction as a negative NOE. For example, consider the spatial arrangement shown in the figure on the left, where protons A-B, and B-C are spatially close, but A-C are far apart. The figure on the right shows a 1D NOESY spectrum when proton A is selectively excited. Since A and B are close, a positive NOE is observed at B. However, an indirect NOE may also appear at C via B. This indirect NOE appears in the opposite direction to the positive NOE. It's important to note that this signal does not indicate that A-C are spatially close. This phenomenon is most likely to occur when the three protons are arranged nearly linearly in space, and it is known as the "three-spin effect."
When, three protons are arranged spatially special, NOE may not be observed even if the distances are close.
As a special case of the three spin effect, let me explain the triangle problem. As shown in the above figure, when three protons are arranged spatially special, this problem can occur.
When A is selectively excited, not only the distance between A and B, but also A and C are close. Thus, positive NOE is supposed to be seen between A and C, the forecasted spectra is the right in the above figure.
However, the positive NOE from A to C shown in orange and the indirect NOE via B shown in green can cancel each other out, resulting in no observable NOE. Theoretically, this cancellation occurs when the ratio of distances between each pair of protons is 1:1:1.26. In other words, even if the protons are spatially close, NOE may not be observed.

Reference

  • Claridge, D.W.T.(2004), *Yūki kagaku no tame no kōbun Kainou NMR tekunikku* [High-resolution NMR techniques for organic chemistry] (T.Takeuchi & M. Nishikawa (Trans.) Kodansha Scientific.
  • Fukushi, E. & Sohmiya H. (2007), *Korenara wakaru nijigen NMR* [Understanding 2D NMR made easy]. Kagaku Dojin.

Structural Analysis of Organic Compound Using 2D - NMR Spectrum

Structural Analysis of Organic Compound Using 2D - NMR Spectrum

In this column, we will explain the structural analysis of organic compound by using 2D-NMR spectrum.

Structural Analysis of one-dimensional NMR (1D-NMR)

One-dimensional NMR (1D-NMR) is the most basic measurement of NMR spectroscopy, in which the spectrum is displayed along a horizontal axis (chemical shift). As introduced in the previous column, information on chemical shift, integration ratio, and splitting pattern (coupling) is used to analyze the structure of an object. However, it can be difficult to perform structural analysis using only 1D-NMR data when the number of detected signals is large, when signals appear in overlapping positions, or when couplings are complex.
The two-dimensional (2D-NMR) spectral analysis method introduced in this column can be used extensively regardless of the object to be analyzed, since the basic sequence of steps is followed. In addition, by following the basic procedures, even beginners can mechanically perform structural analysis.

What is 2D-NMR?

2D-NMR is a method used to check a more detailed molecular structure than 1D-NMR. Since there are many 2D-NMR measurement methods, here, we introduce representative measurement methods.
Measurement Name What you can find out
COSY A method to observe coupling between neighboring same nuclides (frequent:1H/1H)
INADEQUATE A method to observe coupling between neighboring 13Cs
HMQC/HSQC A method to observe coupling of a directly coupled heterogeneous nuclei (frequent: 1H/13C)
HMBC/H2BC A method to observe the correlation signals of heterogeneous nuclei(frequent: 1H/13C) via 2-3 bonds
ADEQUATE A method to detect 13C-13C coupling by 1H observation
NOESY/ROESY A method to observe the correlation signals between 1Hs that are closely located(with NOE interaction)

Structural Analysis Using J Coupling Correlation

This time we introduce an example of structural analysis using "J Coupling Correlation" which is an interaction via chemical bonding. Structural analysis of an object is conducted by following the 4 steps below.
Structural Analysis Using J Coupling Correlation

In Step 1, we use a 1D-NMR spectrum of 13C and DEPT to determine and estimate the atomic group and functional group.

In Step 2, we use HMQC and 1D spectrum of 1H to find 13C and 1H that are directly bonded.

In Step 3, we use COSY to examine and locate which is the neighboring 1Hs.

By using information obtained in Step 2 and 3, substructures can be determined to some extent.

In Step 4, we use HMBC to determine the structure of the remaining parts, by considering the connection of 13C and 1H which are further apart.

Structural analysis example of C6H10O2

Now we will explain the structural analysis of a specific sample.

This time, we use a substance whose molecular formula C6H10O2 is only known and dissolve in heavy chloroform.

Step 1 13C (1D-NMR) & DEPT

In Step 1, we use a 1D-NMR spectrum of 13C and DEPT to determine and estimate atomic group and functional group. First, as shown in Fig. 1, we put a symbol on each signal in 1D-NMR spectrum of 13C. Starting with the signal on the right side of the spectrum, from the high-field side, add symbols A, B, C... and so on. The chemical shift of each signal is read to one decimal place. The read information is summarized in Table 1.

Fig. 1 13C Spectra of C6H10O2
Signal Chemical Shift
A 14.1 ppm
B 17.6 ppm
C 59.8 ppm
D 122.8 ppm
E 144.0 ppm
F 166.2 ppm
Table 1 13C Spectra Information

Next, we will confirm the results of the DEPT spectra. Since DEPT observes the carbon to which hydrogen is directly bonded, the series of the carbon atom of interest (the number of hydrogen directly bonded to the carbon of interest) is known. In other words, it is possible to identify atomic groups such as CH3, CH2, and CH. (DEPT cannot detect spectra of quaternary carbons.) DEPT provides three types of spectra, depending on the measurement parameters (DEPT135, DEPT90, and DEPT45).

Table 2 lists the signal appearance patterns in the three types of DEPT spectra. In DEPT135, the signals of CH3 and CH are detected upward, while the signal of CH2 appears downward (opposite phase). It is often possible to discriminate CH3 or CH from the chemical shift. In many cases, the DEPT135 measurement alone is sufficient. If discrimination by DEPT135 alone is difficult, measure DEPT90, in which only the signal of CH is detected.

Table 2 DEPT Signal Appearance Patterns

We will use the DEPT spectra to determine the atomic groups of the sample. As shown in Figure 2, we will discriminate the atomic groups of each signal by comparing the signal appearance patterns of the 1D-NMR spectrum of 13C and the DEPT spectrum.

-----
DEPT135・・・Upward signal: CH3 or CH, Downward signal: CH2
DEPT90・・・ Detectable signal : CH
Signal detected only in the 1D-NMR spectrum of 13C and not detected in the DEPT spectrum: quaternary carbon
-----

Using the information so far, we were able to determine that the atomic groups of each signal are, from right to left, A (CH3), B (CH3), C (CH2), D (CH), E (CH), and F (quaternary carbon).

Fig. 2 13C Spectra and DEPT Spectra of C6H10O2

The information so far is summarized in Table 3. This information is used to estimate the atomic groups and functional groups.
Signal Chemical Shift Atomic Group
A 14.1 ppm CH 3
B 17.6 ppm CH 3
C 59.8 ppm CH 2
D 122.8 ppm CH
E 144.0 ppm CH
F 166.2 ppm C

Table 3 13C Spectra Information (2)

Figure 3 shows the chemical shift table for 13C. At an approximate border of 100 ppm, the saturated carbon signals are appeared on the right and the unsaturated carbon signals on the left. We will compare the information summarized in Table 3 with the chemical shift table. First, if we look at D (122.8 ppm) and E (144.0 ppm), we see that they are in the region of unsaturated carbon, indicating that this CH is derived from unsaturated carbon. Next, if we look at F, the chemical shift value is 166.2 ppm, which is in the ester region. The molecular formula of this substance is C6H10O2, and since there are two Os, the quaternary carbon of F is presumed to be derived from the COO group.

Figure 3 13C Chemical Shift Table

The information so far is summarized in Table 4.
Signal Chemical Shift Atomic Group
A 14.1 ppm CH3
B 17.6 ppm CH3
C 59.8 ppm CH2
D 122.8 ppm CH=
E 144.0 ppm CH=
F 166.2 ppm COO

Table 4 13C Spectra Information (3)

Step 2 HMQC & 1H

In Step 2, we will use HMQC and 1D-NMR spectra of 1H to find directly bonded 13C and 1H combinations. HMQC will tell you the combination of 1H and 13C that are directly bonded. Spin couplings are written with a symbol, such as 1JCH. The number of bonds is written in the upper left corner of the J representing the spin coupling, and the nucleus to which it is bound is written in the lower right corner of the J.

Figure 4 shows an HMQC spectrum. In a 2D-NMR spectra, a high-resolution 1D-NMR spectrum is displayed at each axis; for HMQC, the 1H spectrum is displayed on the X-axis and the 13C spectrum on the Y-axis. In Step 2, the same symbols that were attached to each signal in the 13C spectrum in Step 1 are attached to the signals in the 13C spectrum on the Y-axis, with the upper side of the Y-axis representing the high-field side (small chemical shift) and the lower side representing the low-field side (large chemical shift).

Next, draw a line from the 13C signal on the Y-axis to the HMQC correlation signal and confirm which 1H is directly bonded to 13C. For example, if we focus on the 13C signal of C, we draw a line toward the side, and when the line hits the correlation signal, we draw a line up from there. The 1H signal that we have reached here is the counterpart directly bonded to signal C of 13C. When the corresponding 1H is found in this way, the 1H is marked with the same symbol as the counterpart 13C. Since it is the counterpart of the signal C of 13C, we will also put the symbol C on this 1H. All correlated combinations will be given a similar symbolization.

Fig.4 HMQC Spectra

Step 3 1H-1H COSY

In Step 3, we will use 1H-1H COSY to see which are the neighboring 1Hs. COSY shows the spin-coupled 1H connections. What can be observed mainly is the correlation between 1Hs via three couplings, as shown in Figure 5. In symbols, we denote this as 3JHH. As shown in Fig. 5, if 1Hs in a 3JHH relationship are lined up, we can follow the 1H connections one after another.

In COSY, in addition to the 3JHH, a 2JHH through two couplings and a remote 4JHH can also be observed.

However, the most important and necessary information is the correlation between 1Hs at the neighboring 13C, 3JHH.

Fig. 5 Correlation of 1Hs via Three Bonds

Fig. 6 shows the actual COSY spectra. COSY shows 1H spectra both on the X-axis and Y-axis.

In Step 1, we assigned a symbol to each 13C signal. In Step 2, we assigned a symbol to each 1H signal that is directly coupled to 13C using HMQC. In Step 3, we first assigned the same symbols to the signals in the 1H spectrum on the X and Y axes as we did in Step 2. We looked at the HMQC spectra we used earlier (Figure 4) and copied the symbols attached to the signals in the 1H spectra on the X-axis as they were.

Note that in Figure 6, the symbols are alphabetically ordered from the high-field side, but this just happened to be the case for this sample. The symbols on the 1H signal are not always in alphabetical order.

Also, in the COSY spectra, signals line up on the diagonal line (called diagonal signals) when the diagonal line is drawn. But these are not used as information for structural analysis. The off-diagonal signals are the COSY correlation signals.

We draw a line from the correlated signal toward the X- and Y-axes spectra and find the 1Hs that are spin-coupled to each other. For example, if we focus on the correlation signal circled in green, we find the 1H signal C on the X-axis and the 1H signal A on the Y-axis. Therefore, we can see that signals C and A are spin-coupled (neighboring) to each other. Once the counterpart is found, we add a symbol to the correlated signal. For example, C/A, to specify that they are 1H signals in a 3JHH relationship. Since the correlated signals of COSY appear in a line symmetric position with respect to the diagonal, we find the 3JHH counterpart in the same way for all correlated signals.

Fig. 6 COSY Spectra

Once the information on the correlated signals of COSY is available, a correlation table for each signal is created.

Table 5 is the correlation table for COSY. Write the symbols for the 1H signals in vertical and horizontal order, as per the spectrum of the data on the 1H-1HCOSY axis. (In the case of this sample, they happen to be arranged alphabetically, but you should look at the COSY spectra and list them in chemical shift order.) For example, if the correlated signals are C and A, mark the intersection of C and A. In this way, all correlated signals are entered in the COSY correlation table.

Table 5 COSY Correlation Table

Fig.7 shows the diagram of correlation signal of COSY.

In this sample, COSY correlation signal was observed between 1H of C and 1H of A.

The HMQC spectra also indicated that 1H of C and 13C of C are directly coupled. In the same way, 1H of A and 13C of A are directly coupled.

Therefore, COSY correlation signals between C and A can induce that 13C of C and 13C of A are neighboring. In other words, using the COSY information, we can indirectly find the neighboring 13C connections from the neighboring 1H connections.

Fig. 7 COSY Correlation Signals

The information that we have so far is the 13C spectrum information (Table 4) and the COSY correlation table (Table 5). From the COSY correlation table, we can see that the carbon atoms derived from the 13C signals A-C, B-E, and D-E are neighboring to each other. Combined with the information from the 13C spectrum, when we rewrite the symbols as atomic groups, we have A-C : "CH3-CH2-", B-E-D : "CH3-CH=CH-", and F : "COO" in Step 1. Therefore, we know that this compound is composed of the following three substructures.

----
Substructure 1:CH3-CH2- ・・・ A-C
Substructure 2: CH3-CH=CH- ・・・ B-E-D
Substructure 3: COO ・・・ F
----

Next, we will consider what kind of molecule these will make by connecting each substructure (Figure 8). We will list all possible molecular structures from the combination of the three substructures. When considering molecular structures, it is easier to focus on the substructure containing CH3 because CH3 is located at the end of the structure. In this case, there are two possible structures. Firstly, Substructures 1 and 2 are found to be the ends of the molecular structure because they contain CH3.And since 1 and 2 are located at the ends, we can predict that substructure 3 is sandwiched between 1 and 2. The different orientations of substructure 3 allowed us to create two different inferred structures (I, II). The orientation of the COO can be used to determine which inferred structure is more reasonable. Therefore, HMBC measurements are performed to confirm the long-range spin coupling with respect to the COO.

Fig. 8 Combination of Three Substructures

Step 4 HMBC

In Step 4, we will consider long-range spin coupling between 13C and 1H by using HMBC.

In HMBC, we can observe the correlations of 2JCH through two bonds or 3JCH through three bonds, as shown in Figure 9.

Fig. 9 CH Correlation through 2 or 3 Bonds

Fig. 10 is the actual spectra of HMBC. The HMBC axes are the same as HMQC's case, 1H spectra on X-axis and 13C spectra on Y-axis. In Step 4, we will put the same symbol that we assigned for HMQC spectra in Step 2, for each signal of a 1D-NMR spectrum on the X and Y axes.

Next, we draw lines from the correlation signals to the X- and Y-axis spectra to find the spin-coupled 13C and 1H combinations. Here, the two horizontally-lined up signals circled in orange are the signals of 1JCH, a direct coupling of 1H and 13C. Since these are remnant signals, they are not observed in all direct couplings. As the direct couplings have already been observed by HMQC, it is not necessary to focus on them here. In the HMBC spectrum, only the long-range information is read, ignoring the direct coupling signal. For example, if we focus on the correlation signal circled in green, we can see that 13C of C and 1H of A are long-range spin coupled to each other, since tracing to Y-axis results in C and to X-axis results in A.

Once the counterpart is known, add a symbol to the correlation signal, e.g., A/C, to indicate that it is a long-range spin-coupled 1H and 13C of 2JCH or 3JCH. For all correlation signals, do the same to find a long-range spin-coupling counterpart and mark the correlation signal with a symbol. Once you have written out the information for the HMBC correlation signals, create a correlation table.

Fig. 10 HMBC Spectra

Table 6 is the correlation table for HMBC. We write the symbol for the 1H signal horizontally and the symbol for the 13C signal vertically, as per the spectrum. (Note that the symbols for the 1H signals are not necessarily in alphabetical order.) Correlation signals for the HMBC spectra will be filled in. For example, if the correlation signal is 13C for C and 1H for A, mark the intersection of C for 13C and A for 1H in the correlation table. In this way, all correlation signals are entered in the HMBC correlation table.

Table 6 HMBC Correlation Table
We use the information in the HMBC to connect the substructures.
Look at the HMBC correlation table (Table 7) that you filled out earlier. In the correlation signal marks, bonds that have already been analyzed by COSY in Step 3 are marked with ○, and bonds that were found for the first time in HMBC are marked with ●. The bond that was found for the first time in HMBC is the COO correlation for F. This indicates that F is a long-range spin couplings with C, D, and E.
Table 7 HMBC Correlation Table (2)

We will look at the long-range correlations of F-C, F-D, and F-E to consider which direction is correct for the COO to be attached, among the inferred structures I and II. (Figure 11).

First, in inferred structure I, F's 13C and C's 1H have three bonds "3JCH", ; F's 13C and D's 1H have two bonds "2JCH", ; and F's 13C and E's 1H have three bonds "3JCH".

Next, let us look at the inferred structure II. The 1H of 13C and C in F is "2JCH" with two bonds, the 1H of 13C and D in F is "3JCH" with three bonds, and the 1H of 13C and E in F is "4JCH" with four bonds.

Since 4JCH is not likely to be observed, we can expect that the inferred structure I is a reasonable structure.

Fig. 11 Comparison between Inferred Structures I and II

Finally, the inferred structure I is checked against the information in the 1D-NMR spectrum of 13C to confirm if it is really correct. In the inferred structure I, CH2 of the symbol C is bonded to oxygen. And the 13C chemical shift of this C was 59.8 ppm (Table 1).

The 13C chemical shift table for CH2 (Table 8) shows that normal CH2 is observed at 20-45 ppm, while CH2 bonded to oxygen is observed at 40-70 ppm with a lower field shift.

From this fact, we can conclude that "Structure I is correct".

Chemical Shift
- CH2 - 20 -45 ppm
- CH2O - 40 - 70 ppm

Table 8 13C Chemical Shift for CH2

How to read NMR spectra from the basics (chemical shift, integration ratio, coupling)

How to read NMR spectra from the basics (chemical shift, integration ratio, coupling)

This column provides easy-to-understand explanations about what we can learn from the NMR spectra (chemical shift, integration ratio, and coupling)

What we can learn from the NMR spectra

There are three main things that we can learn from the NMR spectra.
  1. Horizontal axis (chemical shift):
    The horizontal axis contains information about the type of functional group and molecule conformation. From the position where the spectrum appears (numerical value on the horizontal axis), it is possible to predict what kind of functional group and molecule conformation are contained in the molecule to be measured. 
  2. Integration ratio (signal area ratio):
    By comparing the integral values of each signal, it is possible to compare the number of functional groups contained in a molecule and to obtain information on the mixing ratio of a mixed sample consisting of multiple molecules.
  3. Splitting pattern (coupling):
    The signal is split due to the influence of another nuclear spin existing near the nuclear spin of interest. Figure 1 shows the 1H NMR spectrum of ethanol. The methyl and methylene group signals show that the signal is not a single signal, but is split into multiple signals. Since the splitting pattern of the signal depends on the number and type of other nuclear spins existing nearby, it is possible to predict the substituents contained in the system from the splitting pattern.
Fig.1 1H NMR spectrum of ethanol (CH3CH2OH)

Fig.1 1H NMR spectrum of ethanol (CH 3CH 2OH)

Reasons for causing differences in horizontal axis (chemical shift)

The difference of chemical shift is due to the strength of the magnetic field received (felt) by the nuclear spin we are focusing on.
As shown in Fig 2, depending on the height of the electron density existing near the nuclear spin, the strength of shielding of the magnetic field (the strength of the magnetic field that the nuclear spin receives) varies.
The electron density existing near the nuclear spin depends on the magnitude of the electronegativity of the atoms existing near the nuclear spin of interest.
When an O (oxygen) atom with high electronegativity exists nearby, electrons are attracted by the O atom, the electron density near the nuclear spin of interest decreases, and the magnitude of the magnetic field that the nuclear spin receives becomes greater.
As the electron density near the nuclear spin decreases(shielding becomes lower), the corresponding signal shifts to the left.
Fig. 2 Difference of strength of shielding the magnetic field
Formula of chemical shift

Example of chemical shift table of 1H

Fig. 3 Correlation diagram of typical functional groups and 1HNMR signal positions

Fig. 3 Correlation diagram of typical functional groups and 1HNMR signal positions

Figure 3 shows a correlation diagram of typical functional groups and 1HNMR signal positions. In the NMR spectra, the right side is generally called as the high-field side and the left side as the low-field side. The signal appearing at 0 ppm is the signal of the reference material TMS (tetramethylsilane). The chemical shift value is a numerical value that represents the shift from other signals) So, it is necessary to calibrate the reference point with a reference material such as TMS etc. 1H signals from alkyl chains, such as methyl, methylene, and methine, often appear near 1 ppm. And as mentioned above, 1H signals near alcohol and ether groups with neighboring oxygen atoms and 1H signals derived from amino groups with neighboring nitrogen atoms are detected near 3ppm to 4ppm. The signal appearing near 5 ppm is an alkene-derived 1H signal. Furthermore, 1H signals derived from aromatic rings are observed around 7 ppm, and a signal derived from formyl groups such as aldehydes appears around 9 ppm. Signals derived from carboxyl and phenol groups appear around 11 ppm. The position at which the signal appears allows a rough prediction of the type of functional group.

When performing a structural analysis using NMR, please be careful of heavy water exchange in the case where OH or COH groups are included. In solution NMR, the sample is dissolved in a heavy solvent for measurement. If the solvent to be used is heavy water or heavy methanol, heavy water exchange occurs between the D(2H) in the solvent molecule and the 1H in the OH or COH groups, and the 1H signal from the OH or COH groups may not be observed.

Integration ratio

Fig.4 1HNMR spectrum of benzyl acetate

Fig.4 1HNMR spectrum of benzyl acetate

The following is a brief introduction to the use of integration ratios. Figure 4 shows the structural formula of benzyl acetate and the 1H spectrum. Looking at the molecular structure of benzyl acetate, we can guess that 1H signals would be observed in three areas related with the CH3 group, the CH2 group, and the aromatic group.

Furthermore, a closer look reveals that benzyl acetate has three 1Hs derived from CH3, two 1Hs derived from CH2, and five 1Hs derived from one substituted aromatic CH. The integration ratio of each signal is calculated to be CH3:CH2:CH = 3:2:5, which indicates that the values predicted from the structure and the actual measured values coincide.

It can also be seen that CH3 is shifted to the left from the area where 1H signal derived from CH3 is often observed (around 1 ppm) due to the influence of the neighboring O atoms.

Examples of the use of integration ratios in mixed samples include the following:
  • Relative quantitative evaluation by comparison of integration values of each component
  • Absolute quantitative evaluation using a standard substance of known purity (q-NMR)
  • Calculation of reaction rate by comparison of integration values before and after the reaction
In both examples, it is important to find the signal that is specific to each component and that can be integrated correctly (i.e., not overlapping with other signals).

Coupling and Spin-Spin Coupling Constant "J"

Fig.5 1H NMR spectrum of 2,4 dimethyl pyrimidine

Fig.5 1H NMR spectrum of 2,4 dimethyl pyrimidine

Finally, we introduce couplings. Coupling refers to the interaction between the nuclear spin of interest and another neighboring nuclear spin. In 1D measurements of 1H NMR, the interaction, "coupling" occurs when nuclear spins are in proximity to each other and induces the NMR signal splits. The unit of the splitting width of spin coupling is expressed in Hz. This number is called the spin coupling constant or J-coupling constant (j-value).

Formula for calculating spin coupling constant (J value)
It is also known that the splitting widths have the same j-value when coupled to each other. In the compound in Fig. 5, Ha and Hx are coupled, so the splitting widths of both Ha and Hx have the same value, 6hz. Thus, when a split peak is observed, the j-value information can be used to determine which signals are coupled.

Splitting pattern due to coupling

Splitting pattern due to coupling
Splitting pattern due to coupling

Let us explain a little more about the splitting pattern caused by coupling. An unsplit signal is called a singlet, denoted by the symbol "s"; a two-divided signal is a doublet, denoted by the symbol "d"; a three-divided signal is a triplet, denoted by the symbol "t"; a triplet has a signal strength ratio of 1:2:1, : A signal that splits into four is a quartet, denoted by the symbol "q." The signal strength ratio for a quartet is 1:3:3:1. Signals with five or more segments are multiplets, indicated by the symbol "m".

Using the 1H NMR spectrum of ethanol as an example, we will explain the splitting of the 1H signal. Focusing on the signal derived from the CH3 group around 1ppm, the number of 1H near by CH3 group is 2 (coupled with CH2 group), so it splits into 2+1=3.

Looking at the signal derived from the CH2 group around 3.5 ppm, the number of 1H near by CH2 group is 3 (coupled with CH3 group), which splits into 3+1=4.

Because the OH signal around 5ppm is not coupled to the near by 1H, it does not split and is in the singlet state. Basically, we can see that the signal splits into the "n+1", "n" means the number of nuclei spins positioning around the nuclear spin of interest.

Featured Image 1 - What Role Does Chemical Ionization Play in GC-MS - JEOL USA.png

What Role Does Chemical Ionization Play in GC-MS?

What Role Does Chemical Ionization Play in GC-MS?

Gas chromatography–mass spectrometry, commonly abbreviated as GC-MS, is designed to determine what chemical species are present in a sample. By coupling chromatographic separation with mass-based detection, GC-MS enables analysts to resolve complex mixtures, like environmental samples, petrochemical fractions, or biological extracts, and assign identities to individual volatile and semi-volatile components. In many workflows, GC-MS identification relies on a combination of chromatographic behavior and mass spectral data to effectively distinguish compounds. For known substances, comparison against established reference libraries is often sufficient. The analytical challenges increase, however, when GC-MS is applied to unknown compounds, newly synthesized materials, or trace-level contaminants that fall outside existing databases. Under these conditions, identification hinges on a more fundamental requirement: establishing the intact molecular mass. Chemical ionization, also abbreviated as CI, can provide this critical information through preserving molecular ions, allowing GC-MS to move beyond pattern matching and toward reliable molecular confirmation.

How Chemical Ionization Controls Energy Transfer in GC-MS

Chemical ionization introduces a different ionization environment within the GC-MS ion source, deliberately designed to control how energy is transferred to analyte molecules. Instead of interacting directly with the analyte, ionization occurs through a reagent gas that acts as an intermediary. Methane, isobutane, and ammonia are commonly selected because they form stable, predictable reagent ions under controlled-source conditions.

Ionization begins with the reagent gas itself. Once ionized, the reagent ions establish a low-energy chemical environment inside the ion source. As neutral analyte molecules enter this region, they undergo ion-molecule reactions, such as proton transfer or adduct formation. Because the analyte is ionized indirectly, the amount of internal energy transferred during the process is tightly constrained.

The analytical impact of this controlled energy transfer is immediate. Molecular bonds are largely preserved, fragmentation is minimized, and the resulting GC-MS spectrum is dominated by a clearly defined molecular ion. Rather than producing dense fragmentation patterns, CI concentrates signal intensity into ions that directly reflect molecular weight. For compounds where molecular mass information is difficult to obtain using more energetic ionization methods, such as electron ionization (EI) or CI, CI provides a clear path to molecular confirmation because it can preserve the intact molecular ion rather than promote extensive fragmentation.

Chemical Ionization as a Mass Verification Tool in GC-MS

In GC-MS workflows, CI is applied to confirm molecular weight. Standard GC-MS conditions generate structurally informative fragmentation patterns, but the energetic nature of these processes can suppress the molecular ion, particularly for thermally labile or highly substituted compounds. Without a clearly observable molecular ion, identification becomes more uncertain and increasingly reliant on indirect evidence.

Chemical ionization addresses such a limitation by preserving molecular integrity during ion formation. Rather than distributing ion signal across numerous fragment ions, the technique concentrates intensity into a small number of species that directly reflect molecular mass, most commonly a protonated molecule or a predictable adduct. This focused spectral output provides clear molecular weight information, especially when reference library matching alone is insufficient.

Beyond molecular weight confirmation, CI also enhances interpretability in demanding GC-MS applications. Cleaner spectra reduce ambiguity during data processing, support more robust spectral deconvolution, and improve discrimination between closely related compounds. For samples containing coeluting components or structurally similar species, this clarity strengthens identification decisions and reduces the likelihood of misassignment.

Enabling Accurate Mass and Elemental Formula Determination

High-resolution GC-MS places additional demands on ionization strategy, as accurate mass measurements are only meaningful when the correct molecular ion is first established. Without clear molecular ion information, even high mass accuracy cannot resolve structural uncertainty. Chemical ionization preserves the molecular ion, allowing the correct nominal mass to be established before exact mass calculations are applied.

With the molecular ion confirmed, mass data can be interpreted with confidence to evaluate plausible elemental compositions. Subtle mass differences can become analytically significant, enabling discrimination between closely related formulas such as CxHyNz variants that may otherwise appear indistinguishable. This capability is particularly important in research, regulatory, and forensic contexts, where compounds may be novel, intentionally modified, or designed to resemble known substances.

By supporting reliable formulate determination, CI extends the role of GC-MS beyond library-based identification to definitive molecular characterization and increases the confidence of analytical conclusions for compounds that cannot be resolved by spectral matching alone.

Selective Detection Using Negative Chemical Ionization in GC-MS

Chemical ionization also enables a highly selective GC-MS approach known as negative chemical ionization, or NCI. In NCI, the ion source favors electron capture over proton transfer, allowing only compounds with sufficient electron affinity to form stable negative ions. This selectivity is governed by chemical structure rather than broad ionization efficiency. Halogenated and nitro-containing compounds respond strongly to NCI, while most background species remain neutral. As a result, NCI GC-MS delivers high sensitivity with minimal chemical noise, which is particularly advantageous for trace-level analysis in complex sample matrices.

Applying Chemical Ionization with JEOL USA

    Featured Image 5 - Overcoming Structural Biology Challenges with Transmission Electron Microscopy (TEM).png

    Overcoming Structural Biology Challenges with Transmission Electron Microscopy (TEM)

    Structural biology aims to understand how biological molecules function by determining their three-dimensional structures. However, many targets are too flexible, unstable, or complex for traditional techniques like X-ray crystallography. These limitations can slow or even halt progress.

    Transmission electron microscopy (TEM) offers powerful solutions to these challenges. Within structural biology, TEM encompasses a range of imaging modes—including cryogenic electron microscopy (cryo-EM), negative staining, and cellular tomography—that together enable researchers to visualize macromolecules, viruses, organelles, and cellular structures at nanometer to near-atomic resolution.

    Cryo-EM, in particular, has become a transformative tool by allowing structural biologists to visualize individual particles in frozen, hydrated states. As a cryogenic mode of TEM, it avoids the need for crystallization or staining and supports high-resolution reconstruction of macromolecular assemblies in near-native conformations. Whether imaging purified molecules or cellular sections, TEM technologies provide essential insights into molecular architecture and function.

    Structural Biology Challenges Addressed by TEM (Including Cryo-EM)

    1. Targets That Resist Crystallization


    The Challenge:
    Many proteins, especially membrane-associated or flexible ones, cannot be crystallized for X-ray analysis.

    TEM Solution: Cryo-TEM allows researchers to image these molecules directly in frozen, hydrated form—bypassing the need for crystallization and enabling structure determination of otherwise elusive targets.Vivamus sagittis lacus vel augue laoreet rutrum faucibus dolor auctor. Duis mollis, est non commodo luctus.

    2. Structural Heterogeneity and Molecular Flexibility


    The Challenge: Many macromolecules exist in multiple conformations, but traditional methods that average all particles together can obscure these distinct structural states.

    TEM Solution: In cryo-EM, sophisticated computational sorting can separate images of particles into different conformational states, allowing individual reconstructions of each. This reveals how molecules change during function.Vivamus sagittis lacus vel augue laoreet rutrum faucibus dolor auctor. Duis mollis, est non commodo luctus.

    3. Large and Asymmetric Assemblies


    The Challenge: Massive molecular complexes often defy symmetry assumptions or size limits used in traditional methods.

    TEM Solution: Cryo-EM handles large complexes—often in the megadalton range—and can reconstruct asymmetric or irregular structures with high detail, enabling study of entire molecular machines.

    4. Capturing Transient or Rare States


    The Challenge: Short-lived intermediates or low-population species are easily missed in ensemble measurements.

    TEM Solution: High-throughput data collection and particle classification in cryo-EM allow detection and reconstruction of rare or transient molecular states critical for understanding dynamic processes.

    5. Structural Context Within Cells and Tissues


    The Challenge: Understanding molecular function requires knowledge of spatial organization in situ.

    TEM Solution: By imaging ultrathin sections of cells or tissues, TEM reveals how macromolecules, organelles, and complexes are arranged within their native cellular environment—essential for linking molecular structure to biological function.

    6. Sample Screening and Quality Control


    The Challenge: Poor sample quality can derail high-resolution studies.

    TEM Solution: Negative-stain TEM offers rapid, low-resolution imaging to assess particle homogeneity, concentration, and integrity before advancing to cryo-EM, streamlining workflow and improving data quality.

    7. Morphological Analysis of Viruses and Organelles


    The Challenge:
    Conventional techniques lack the resolution to resolve subcellular structures and viral architecture.

    TEM Solution: Both conventional and cryo-TEM provide high-resolution views of viruses, vesicles, and organelles, revealing structural details that guide functional and biochemical hypotheses.

    Common Advantages of TEM Techniques in Structural Biology

    Preservation of Sensitive and Unstable Samples

    Cryogenic preparation—used in cryo-TEM—preserves structural integrity by vitrifying samples rapidly in ice. This minimizes damage from vacuum and the electron beam, allowing imaging of fragile biological material.

    Imaging in Native or Near-Native ConditionsHeading

    By avoiding fixation, dehydration, and staining, cryo-TEM captures hydrated samples in near-native states. This helps preserve molecular conformations that more closely reflect their true biological roles.

    Advancing Structural Biology with JEOL

    Featured Image - Is Personalized Structural Biology a new frontier in medicine - JEOL USA - Draft 2.png

    Is personalized structural biology a new frontier in medicine?

    Is personalized structural biology a new frontier in medicine?

      Personalized structural biology is opening new possibilities in how disease is studied and treated. By connecting genetic variation to the three-dimensional structures of proteins, the field offers a clearer view of how molecular changes influence health at an individual level. This added layer of precision allows researchers to see not just where mutations occur, but how they alter the shape and behavior of key biological molecules. As structural insights become more detailed and widely available, their part in guiding diagnostics and treatment is becoming more significant. Advances in structural modeling and patient-specific analysis within personalized structural biology point to a steady and meaningful shift toward a more individualized and structurally informed approach to medicine.

      Turning Molecular Insight into Medical Precision

      Structural biology has long helped scientists understand the shapes and functions of proteins, often relying on shared or average genetic sequences to construct generalized models. These models have advanced scientific understanding but rarely capture the variation present in individual patients. In medicine, that diversity can make all the difference.

      Personalized structural biology brings that difference into focus. Rather than examining mutations in abstract, researchers can map them onto the actual contours of a patient’s proteins. They can also observe how those changes affect folding, flexibility, or molecular interactions. Such a level of detail produces a more direct link between genetic variation and clinical outcome, helping to refine both diagnosis and treatment with a degree of precision that standard models cannot provide.

      Why Personalized Structural Biology Represents a New Frontier in Medicine

      From Genetic Code to Functional Clarity

      Genomic sequencing identifies where mutations occur, but their consequences are often unclear. Through leveraging patient-specific models, personalized structural biology reveals how mutations alter protein structure and function, bringing their molecular consequences into sharper focus. Subtle changes in folding or binding can interrupt biological function or influence how a person responds to treatment. Clarifying the structural impact of each variant turns genetic data into a practical tool for guiding care, marking a step toward more precise and individualized medicine.

      Tailoring Therapies to Molecular Architecture

      Diseases such as cancer, neurodegenerative disorders, and inherited metabolic conditions often arise from subtle alterations in protein structure. Personalized structural biology allows drug developers to account for these nuances. Whether designing inhibitors that target specific mutant conformations or modifying biologics to better engage dysfunctional proteins, treatments could potentially be matched to the molecular signature of every patient.

      Anticipating and Overcoming Drug Resistance

      Resistance often emerges when mutations change a drug’s binding site on its target protein. Structural modeling allows clinicians to predict which mutations may cause resistance and identify alternative therapeutic strategies. In oncology, for example, this capability is increasingly applied to design second-line therapies before resistance arises. With personalized structural biology integrated into said process, the ability to tailor medical treatments to emerging molecular changes becomes a more practical part of care.

      Building the Foundation for Virtual Patients

      Personalized structural biology is shaping a new dimension of patient care through the development of digital twins. Integrating structural models with other layers of omics data allows researchers to build computational representations of individual patients. Each model captures the molecular details that influence disease progression and drug response, generating a platform to explore and adjust treatments in a virtual setting. This shift toward simulation-based medicine reflects the growing influence of structural insight in advancing more personalized and predictive healthcare.

      Key Techniques Powering Tailored Insights

      The progress of personalized structural biology in medicine depends on tools that can accurately visualize proteins at the molecular level. Three techniques in particular, cryo-electron microscopy (cryo-EM), X-ray crystallography, and nuclear magnetic resonance (NMR) spectroscopy, have been instrumental when used for personalized structural biology, connecting genetic variation to structural and functional outcomes.

      Cryo-EM produces high-resolution images of proteins in conditions that closely resemble their natural environment. It excels at studying large and flexible protein complexes, which are often implicated in cancer and infectious diseases. X-ray crystallography on the other hand can be used to decipher structures of smaller proteins or protein domains, and in combination with patient-derived sequences, both cryo-EM and X-ray crystallography help reveal how specific mutations reshape binding pockets or disrupt key structural features. These insights are essential for designing therapies that align with the molecular characteristics of individual patients.

      NMR spectroscopy complements this through its capacity to capture the movements and flexibility of proteins in solution. This method proves especially valuable in diseases where structural instability plays a central role, such as ALS or Alzheimer's. NMR also contributes to drug discovery by analyzing how small molecules bind to mutant proteins under physiologically relevant conditions.

      Together, cryo-EM, X-ray crystallography, and NMR enhance the capabilities of personalized structural biology in medicine. Their combined strengths deepen insight into protein variation and support the integration of structural information into clinical workflows, advancing a more predictive and patient-tailored approach to medical care.

      Enhancing Precision Medicine Through Personalized Structural Biology

      Personalized structural biology is emerging as a transformative force in medicine, offering a way to connect genetic differences with the structural changes that drive disease. The increasing focus on protein-level insight calls for technologies capable of capturing structural detail with exceptional clarity. JEOL USA’s cryo-electron microscopes and NMR spectroscopy systems provide the advanced performance needed to support the growing demands of precision medicine. Visit our website to see how our CRYO ARM series and high-field NMR systems can elevate your personalized structural biology research with precision and reliability.

      Cryo-EM: Integrating Structural Bioinformatics for Functional Insights

      Cryo-EM: Integrating Structural Bioinformatics for Functional Insights

      Summary

      Cryo-electron microscopy (cryo-EM) determines the three-dimensional structures of biological macromolecules by transforming raw electron micrographs into detailed density maps. While these maps provide crucial information about molecular architecture and spatial organization, they require further interpretation to extract biological function. Researchers employ computational modeling tools to construct atomic-resolution structures from the density data, which structural bioinformatics tools then analyze to identify functionally relevant features. This powerful combination of cryo-EM and bioinformatics creates an integrated pipeline encompassing everything from initial image processing to final biological interpretation. The approach proves particularly valuable for studying complex, flexible molecular assemblies, where structural bioinformatics plays a critical role in converting structural data into meaningful mechanistic understanding.

      Why Structural Bioinformatics Integration Enhances Cryo-EM

      As cryo-EM workflows have grown more sophisticated, researchers are now generating increasingly large and complex datasets. However, without structural bioinformatics, much of this data remains underexploited. Integrating structural bioinformatics aligns model building with biological interpretation, enabling more accurate atomic models and stronger connections to functional insights. It also introduces essential validation steps that minimize the risk of overfitting and maximize the correctness of structural conclusions. Incorporating structural bioinformatics into the cryo-EM pipeline ensures that the growing volume of data not only yields higher-quality models but also enhances our understanding of biological mechanisms.

      Preparing Cryo-EM Data for Structural Interpretation

        The Cryo-EM pipeline involves processing thousands of low-contrast, two-dimensional images obtained from vitrified biological samples in various orientations. These raw projections undergo frame alignment, motion correction, and contrast transfer function (CTF) estimation to minimize noise and enhance signal quality. Next, individual particle images are identified, extracted, and subjected to classification and averaging, improving signal across different views. This multi-step refinement process ultimately generates a three-dimensional density map, revealing the overall architecture and spatial organization of the macromolecule.

        While the density map provides a structural framework, it often lacks the atomic-level detail necessary to fully elucidate molecular function given that most of these maps are at resolutions not better than 1.22Å - the threshold for claiming true atomic resolution.. To bridge this gap, further modeling and refinements are performed to interpret the map at higher resolution and establish meaningful connections between structure and biological mechanism.

        Interpreting Cryo-EM Maps with Computational Tools

        Cryo-EM density maps serve as the foundation for determining atomic models of biological macromolecules through a multi-stage computational pipeline. The process begins by fitting available structural templates into the density using rigid-body alignment or, for novel structures without templates, generating de novo atomic models that match the experimental density. Often primary sequence information is utilized at bulky side chains that can provide for convenient landmarks at this step. These initial models then undergo iterative refinements to simultaneously optimize two critical parameters: the agreement with experimental density and maintenance of proper stereochemical constraints. The refinement process is accompanied by comprehensive validation protocols including geometric checks of bond lengths and angles, quantitative evaluation of local map-model correlation, and verification of sequence register and side chain conformations. This rigorous computational transformation from raw density maps to validated atomic models enables researchers to reliably interpret structural features and derive mechanistic insights, particularly for large, dynamic complexes that pose challenges for conventional structural biology approaches. The integration of advanced modeling algorithms with experimental cryo-EM data has established an essential workflow for bridging structural determination and biological function.

        Supporting Complex Structural Analysis

        Studying large molecular assemblies, flexible proteins, or multi-conformational systems requires more than high-resolution imaging—it demands computational tools capable of resolving structural heterogeneity. Computational cryo-EM techniques separate distinct conformational states from mixed populations, and build upon these results to generate refined atomic models for each state. This approach reveals how individual conformations contribute to molecular function, enabling researchers to investigate dynamic behavior, map conformation-specific binding sites, and correlate structural transitions with biological activity. Such analyses are indispensable for therapeutic design, viral mechanism elucidation, and understanding the operational principles of molecular machines.

        Linking Structure to Function with Structural Bioinformatics

        A validated atomic model represents just the first step toward biological understanding. Structural bioinformatics provides the critical link between this model and its functional context by leveraging curated sequence databases and computational tools. Through comparative analysis, researchers can identify functional domains, conserved motifs, and evolutionary relationships that may not be immediately apparent from structural data alone. These annotations help to interpret the molecular model by revealing potential binding sites, allosteric regions, and mechanistic clues about the molecule’s role in cellular processes.

        The integration of cryo-EM-derived structures with structural bioinformatics transforms atomic coordinates into functional hypotheses. This synthesis enables researchers to propose testable mechanisms, rationalize disease-associated mutations, or identify targets for therapeutic intervention. By bridging structural data with biological knowledge, the workflow culminates in a dynamic, functionally annotated model—one that supports deeper mechanistic studies and accelerates translation from structure to biological insight.

        Data Quality and Instrumentation

        The reliability of computational cryo-EM techniques depend on high-quality image data. Each step—from motion correction and particle alignment to 3D reconstruction—requires consistent, high-contrast input. While advanced algorithms can refine imperfect data, they cannot overcome fundamental limitations imposed by poor imaging conditions. Robust instrumentation and optimized sample preparation form the bedrock of successful structural determination, ensuring that downstream analyses yield biologically meaningful interpretations.

        Powering Integrated Structural Workflows with JEOL USA

        The synergy between cryo-EM instrumentation and computational tools is critical for transforming raw data into biological insight. JEOL’s cryo-EM systems, the CRYO ARM™ 200 and CRYO ARM™ 300, are engineered to meet these demands. Featuring cold field emission sources, in-column energy filters, and automated specimen handling, these platforms deliver the stability and resolution required for high-fidelity data collection.

        For researchers pursuing dynamic or challenging targets, JEOL’s solutions provide the technical foundation to bridge structural biology and mechanistic discovery. Contact our experts to explore how our cryo-EM technology can elevate your research.

        © Copyright 2025 by JEOL USA, Inc.
        Terms of Use
        |
        Privacy Policy
        |
        Cookie Preferences