Monday, December 14, 2020

Version 3.0 of QSoas is out

After almost two years of development, version 3.0 of QSoas is finally out ! It brings in a number of new features.

An expert mode for fitting

Undoubtedly the most important feature in the new version is a complete upgrade of the fit system, which now features an expert mode, turned on by using the /expert=true option with the fit commands. The expert mode features a command prompt that looks like the normal command prompt, in which it is possible:
  • to fix/unfix, set, reset and load ans save parameters;
  • export parameters or computed datasets;
  • launch the fit;
  • select fit engines and configure them;
  • list of all the fits that were run in the session with their starting and finishing parameters (see picture), and reuse them;
  • run scripts;
  • fit data completely non-interactively;
  • and, last but not least, perform parameter space exploration, which consists in running (automatically) a number of fits with different (random) starting parameters to maximize the chances of finding out the best parameters for the fit, avoiding local minima.
The latter feature is very important when running fits with many parameters. In that case, there are a number of local minima, and it is necessary to try a number of different starting parameters to really find the best parameters. The new parameter space exploration feature makes it much easier than before, with an interface that allows easily finding the best parameters tried so far and reuse them.

A new documentation system

Another very important update is the inclusion of a new, offline, documentation system. The documentation features browsing via table of contents, a command index and text search. It also features the possibility to copy commands from the help to the command prompt, or even run them directly. To top it all, it comes with a series of startup tips that might teach you a thing or two about QSoas (try hitting Show random to learn new tricks !).

Many other features

For the full list of changes, please see the changelog. Apart from the changes described above, these are my favorites:
  • running the command from the menu now gives a dialog box for selecting all the arguments and options, which is very useful to discover the possibilities of the various commands;
  • features that make the statistics and meta-data much easier to use with commands like apply-formula, like automatic code completion;
  • a new system for permanently storing meta-data using record-meta;
  • a new fit for fitting implicit equations, in which one cannot express \(y = f(x)\), but only \(f(x,y)=0\);
  • and, of course, a brand shiny new logo by Marta Meneghello !

To get the new version, you can just download the source code from the downloads page, where you can also purchase precompiled versions for Windows and MacOS. You can also clone the source from the GitHub repository.

About QSoas

QSoas is a powerful open source data analysis program that focuses on flexibility and powerful fitting capacities. It is released under the GNU General Public License. It is described in Fourmond, Anal. Chem., 2016, 88 (10), pp 5050–5052. Current version is 3.0. You can download its source code there (or clone from the GitHub repository) and compile it yourself, or buy precompiled versions for MacOS and Windows there.

Monday, November 23, 2020

QSoas tips and tricks: using meta-data, first level

By essence, QSoas works with \(y = f(x)\) datasets. However, in practice, when working with experimental data (or data generated from simulations), one has often more than one experimental parameter (\(x\)). For instance, one could record series of spectra (\(A = f(\lambda)\)) for different pH values, so that the absorbance is in fact a function of both the pH and \(\lambda\). QSoas has different ways to deal with such situations, and we'll describe one today, using meta-data.

Setting meta-data

Meta-data are simply series of name/values attached to a dataset. It can be numbers, dates or just text. Some of these are automatically detected from certain type of data files (but that is the topic for another day). The simplest way to set meta-data is to use the set-meta command:
QSoas> set-meta pH 7.5
This command sets the meta-data pH to the value 7.5. Keep in mind that QSoas does not know anything about the meaning of the meta-data[1]. It can keep track of the meta-data you give, and manipulate them, but it will not interpret them for you. You can set several meta-data by repeating calls to set-meta, and you can display the meta-data attached to a dataset using the command show. Here is an example:
QSoas> generate-buffer 0 10
QSoas> set-meta pH 7.5
QSoas> set-meta sample "My sample"
QSoas> show 0
Dataset generated.dat: 2 cols, 1000 rows, 1 segments, #0
Flags: 
Meta-data:	pH =	 7.5	sample =	 My sample
Note here the use of quotes around My sample since there is a space inside the value.

Using meta-data

There are many ways to use meta-data in QSoas. In this post, we will discuss just one: using meta-data in the output file. The output file can collect data from several commands, like peak data, statistics and so on. For instance, each time the command 1 is run, a line with the information about the largest peak of the current dataset is written to the output file. It is possible to automatically add meta-data to those lines by using the /meta= option of the output command. Just listing the names of the meta-data will add them to each line of the output file.

As a full example, we'll see how one can take advantage of meta-data to determine the position of the peak of the function \(x^2 \exp (-a\,x)\) depends on \(a\). For that, we first create a script that generates the function for a certain value of \(a\), sets the meta-data a to the corresponding value, and find the peak. Let's call this file do-one.cmds (all the script files can be found in the GitHub repository):

generate-buffer 0 20 x**2*exp(-x*${1})
set-meta a ${1}
1 
This script takes a single argument, the value of \(a\), generates the appropriate dataset, sets the meta-data a and writes the data about the largest (and only in this case) peak to the output file. Let's now run this script with 1 as an argument:
QSoas> @ do-one.cmds 1
This command generates a file out.dat containing the following data:
## buffer       what    x       y       index   width   left_width      right_width     area
generated.dat   max     2.002002002     0.541340590883  100     3.4034034034    1.24124124124   2.162162162161.99999908761
This gives various information about the peak found: the name of the dataset it was found in, whether it's a maximum or minimum, the x and y positions of the peak, the index in the file, the widths of the peak and its area. We are interested here mainly in the x position.

Then, we just run this script for several values of \(a\) using run-for-each, and in particular the option /range-type=lin that makes it interpret values like 0.5..5:80 as 80 values evenly spread between 0.5 and 5. The script is called run-all.cmds:

output peaks.dat /overwrite=true /meta=a
run-for-each do-one.cmds /range-type=lin 0.5..5:80
V all /style=red-to-blue
The first line sets up the output to the output file peaks.dat. The option /meta=a makes sure the meta a is added to each line of the output file, and /overwrite=true make sure the file is overwritten just before the first data is written to it, in order to avoid accumulating the results of different runs of the script. The last line just displays all the curves with a color gradient. It looks like this:
Running this script (with @ run-all.cmds) creates a new file peaks.dat, whose first line looks like this:
## buffer       what    x       y       index   width   left_width      right_width     area    a
The column x (the 3rd) contains the position of the peaks, and the column a (the 10th) contains the meta a (this column wasn't present in the output we described above, because we had not used yet the output /meta=a command). Therefore, to load the peak position as a function of a, one has just to run:
QSoas> load peaks.dat /columns=10,3
This looks like this:
Et voilà !

To train further, you can:

  • improve the resolution in x;
  • improve the resolution in y;
  • plot the magnitude of the peak;
  • extend the range;
  • derive the analytical formula for the position of the peak and verify it !

[1] this is not exactly true. For instance, some commands like unwrap interpret the sr meta-data as a voltammetric scan rate if it is present. But this is the exception.

About QSoas

QSoas is a powerful open source data analysis program that focuses on flexibility and powerful fitting capacities. It is released under the GNU General Public License. It is described in Fourmond, Anal. Chem., 2016, 88 (10), pp 5050–5052. Current version is 2.2. You can download its source code there (or clone from the GitHub repository) and compile it yourself, or buy precompiled versions for MacOS and Windows there.

Wednesday, November 11, 2020

Solution for QSoas quiz #1: averaging spectra

This post describes the solution to the Quiz #1, based on the files found there. The point is to produce both the average and the standard deviation of a series of spectra. Below is how the final averaged spectra shoud look like:
I will present here two different solutions.

Solution 1: using the definition of standard deviation

There is a simple solution using the definition of the standard deviation: $$\sigma_y = \sqrt{<y^2> - {<y>}^2}$$ in which \(<y^2>\) is the average of \(y^2\) (and so on). So the simplest solution is to construct datasets with an additional column that would contain \(y^2\), average these columns, and replace the average with the above formula. For that, we need first a companion script that loads a single data file and adds a column with \(y^2\). Let's call this script load-one.cmds:
load ${1}
apply-formula y2=y**2 /extra-columns=1
flag /flags=processed
When this script is run with the name of a spectrum file as argument, it loads it (replaces ${1} by the first argument, the file name), adds a column y2 containing the square of the y column, and flag it with the processed flag. This is not absolutely necessary, but it makes it much easier to refer to all the spectra when they are processed. Then to process all the spectra, one just has to run the following commands:
run-for-each load-one.cmds Spectrum-1.dat Spectrum-2.dat Spectrum-3.dat
average flagged:processed
apply-formula y2=(y2-y**2)**0.5
dataset-options /yerrors=y2
The run-for-each command runs the load-one.cmds script for all the spectra (one could also have used Spectra-*.dat to not have to give all the file names). Then, the average averages the values of the columns over all the datasets. To be clear, it finds all the values that have the same X (or very close X values) and average them, column by column. The result of this command is therefore a dataset with the average of the original \(y\) data as y column and the average of the original \(y^2\) data as y2 column. So now, the only thing left to do is to use the above equation, which is done by the apply-formula code. The last command, dataset-options, is not absolutely necessary but it signals to QSoas that the standard error of the y column should be found in the y2 column. This is now available as script method-one.cmds in the git repository.

Solution 2: use QSoas's knowledge of standard deviation

The other method is a little more involved but it demonstrates a good approach to problem solving with QSoas. The starting point is that, in apply-formula, the value $stats.y_stddev corresponds to the standard deviation of the whole y column... Loading the spectra yields just a series of x,y datasets. We can contract them into a single dataset with one x column and several y columns:
load Spectrum-*.dat /flags=spectra
contract flagged:spectra
After these commands, the current dataset contains data in the form of:
lambda1	a1_1	a1_2	a1_3
lambda2	a2_1	a2_2	a2_3
...
in which the ai_1 come from the first file, ai_2 the second and so on. We need to use transpose to transform that dataset into:
0	a1_1	a2_1	...
1	a1_2	a2_2	...
2	a1_3	a2_3	...
In this dataset, values of the absorbance for the same wavelength for each dataset is now stored in columns. The next step is just to use expand to obtain a series of datasets with the same x column and a single y column (each corresponding to a different wavelength in the original data). The game is now to replace these datasets with something that looks like:
0	a_average
1	a_stddev
For that, one takes advantage of the $stats.y_average and $stats.y_stddev values in apply-formula, together with the i special variable that represents the index of the point:
apply-formula "if i == 0; then y=$stats.y_average; end; if i == 1; then y=$stats.y_stddev; end"
strip-if i>1
Then, all that is left is to apply this to all the datasets created by expand, which can be just made using run-for-datasets, and then, we reverse the splitting by using contract and transpose ! In summary, this looks like this. We need two files. The first, process-one.cmds contains the following code:
apply-formula "if i == 0; then y=$stats.y_average; end; if i == 1; then y=$stats.y_stddev; end"
strip-if i>1
flag /flags=processed
The main file, method-two.cmds looks like this:
load Spectrum-*.dat /flags=spectra
contract flagged:spectra
transpose
expand /flags=tmp
run-for-datasets process-one.cmds flagged:tmp
contract flagged:processed
transpose
dataset-options /yerrors=y2
Note some of the code above can be greatly simplified using new features present in the upcoming 3.0 version, but that is the topic for another post.

About QSoas

QSoas is a powerful open source data analysis program that focuses on flexibility and powerful fitting capacities. It is released under the GNU General Public License. It is described in Fourmond, Anal. Chem., 2016, 88 (10), pp 5050–5052. Current version is 2.2. You can download its source code and compile it yourself or buy precompiled versions for MacOS and Windows there.

Thursday, October 22, 2020

QSoas tips and tricks: generating smooth curves from a fit

Often, one would want to generate smooth data from a fit over a small number of data points. For an example, take the data in the following file. It contains (fake) experimental data points that obey to Michaelis-Menten kinetics: $$v = \frac{v_m}{1 + K_m/s}$$ in which \(v\) is the measured rate (the y values of the data), \(s\) the concentration of substrate (the x values of the data), \(v_m\) the maximal rate and \(K_m\) the Michaelis constant. To fit this equation to the data, just use the fit-arb fit:
QSoas> l michaelis.dat
QSoas> fit-arb vm/(1+km/x)
After running the fit, the window should look like this:
Now, with the fit, we have reasonable values for \(v_m\) (vm) and \(K_m\) (km). But, for publication, one would want to generate "smooth" curve going through the lines... Saving the curve from "Data.../Save all" doesn't help, since the data has as many points as the original data and looks very "jaggy" (like on the screenshot above)... So one needs a curve with more data points.

Maybe the most natural solution is simply to use generate-buffer together with apply-formula using the formula and the values of km and vm obtained from the fit, like:

QSoas> generate-buffer 0 20
QSoas> apply-formula y=3.51742/(1+3.69767/x)
By default, generate-buffer generate 1000 evenly spaced x values, but you can change their number using the /samples option. The two above commands can be combined to just one call to generate-buffer:
QSoas> generate-buffer 0 20 3.51742/(1+3.69767/x)
This works, but it is quite cumbersome and it is not going to work well for complex formulas or the results of differential equations or kinetic systems...

This is why to each fit- command corresponds a sim- command that computes the result of the fit using a "saved parameters" file (here, michaelis.params, but you can also save it yourself) and buffers as "models" for X values:

QSoas> generate-buffer 0 20
QSoas> sim-arb vm/(1+km/x) michaelis.params 0
This strategy works with every single fit ! As an added benefit, you even get the fit parameters as meta-data, which are displayed by the show command:
QSoas> show 0
Dataset generated_fit_arb.dat: 2 cols, 1000 rows, 1 segments, #0
Flags: 
Meta-data:	commands =	 sim-arb vm/(1+km/x) michaelis.params 0	fit =	 arb (formula: vm/(1+km/x))	km =	 3.69767
	vm =	 3.5174
They also get saved as comments if you save the data.

Important note: the sim-arb command will be available only in the 3.0 release, although you can already enjoy it if you use the github version.

About QSoas

QSoas is a powerful open source data analysis program that focuses on flexibility and powerful fitting capacities. It is released under the GNU General Public License. It is described in Fourmond, Anal. Chem., 2016, 88 (10), pp 5050–5052. Current version is 2.2. You can download its source code and compile it yourself or buy precompiled versions for MacOS and Windows there.

Wednesday, October 7, 2020

QSoas quiz #1 : averaging spectra

Here is the first QSoas quiz ! I recently measured several identical spectra in a row to evaluate the noise of the setup, and so I wanted to average all the spectra and also determine the standard deviation in the absorbances. Averaging the spectra can simply be done taking advantage of the average command:
QSoas> load Spectrum*.dat /flags=spectra
QSoas> average flagged:spectra
However, average does not provide means to make standard deviations, it just takes the average of all but the X column. I wanted to add this feature, but I realized there are already at least two distinct ways to do that...

Quiz

Your task is to determine the average and standard deviations of the three spectra located there, (Spectrum-1.dat, Spectrum-2.dat and Spectrum-3.dat). There are at least two ways:
  • One that relies simply on average and on apply-formula, and which requires that you remember how to compute standard deviations.
  • One that is a little more involved, that requires more data manipulation (take a look at contract for instance) and relies on the fact that you can use statistics in apply-formula (and in particular you can use y_stddev to refer to the standard deviation of \(y\)), but which does not require you to know exactly how to compute standard deviations.
To help you, I've added the result in Average.dat. The figure below shows a zoom on the data superimposed to the average (bonus points to find how to display this light red area that corresponds to the standard deviation !).
I will post the answer later. In the meantime, feel free to post your own solutions or attempts, hacks, and so on !

About QSoas

QSoas is a powerful open source data analysis program that focuses on flexibility and powerful fitting capacities. It is released under the GNU General Public License. It is described in Fourmond, Anal. Chem., 2016, 88 (10), pp 5050–5052. Current version is 2.2. You can download its source code or buy precompiled versions for MacOS and Windows there.

Wednesday, September 23, 2020

Tutorial: analyze Km data of CODHs

This is the first post of a series in which we will provide the readers with simple tutorial approaches to reproduce the data analysis of some of our published papers. All our data analysis is performed using QSoas. Today, we will show you how to analyze the experiments we used to characterize the behaviour of an enzyme, the Nickel-Iron CO dehydrogenase IV from Carboxytothermus hydrogenoformans. The experiments we analyze here are described in much more details in the original publication, Domnik et al, Angewandte Chemie, 2017. The only things you need to know for now are the following:
  • the data is the evolution of the current over time given by the absorbed enzyme;
  • at a given point, a certain amount of CO (50 µM) is injected, and its concentration decreases exponentially over time;
  • the enzyme has a Michaelis-Menten behaviour with respect to CO.
This means that we expect a response of the type: $$i(t) = \frac{i_m}{1 + \frac{K_m}{[\mathrm{CO}](t)}}$$ in which $$[\mathrm{CO}](t) = \begin{cases} 0, & \text{for } t < t_0 \\ C_0 \exp \frac{t_0 - t}{\tau}, & \text{for }t\geq t_0 %> \end{cases}$$

To begin this tutorial, first download the files from the github repository (direct links: data, parameter file and ruby script). Start QSoas, go to the directory where you saved the files, load the data file, and remove spikes in the data using the following commands:

QSoas> cd
QSoas> l Km-CODH-IV.dat
QSoas> R
First fit
Then, to fit the above equation to the data, the simplest is to take advantage of the time-dependent parameters features of QSoas. Run simply:
QSoas> fit-arb im/(1+km/s) /with=s:1,exp
This simply launches the fit interface to fit the exact equations above. The im/(1+km/s) is simply the translation of the Michaelis-Menten equation above, and the /with=s:1,exp specifies that s is the result of the sum of 1 exponential like for the definition of above. Then, load the Km-CODH-IV.params parameter files (using the "Parameters.../Load from file" action at the bottom, or the Ctrl+L keyboard shortcut). Your window should now look like this:
To fit the data, just hit the "Fit" button ! (or Ctrl+F).
Including an offset
The fit is not bad, but not perfect. In particular, it is easy to see why: the current predicted by the fit goes to 0 at large times, but the actual current is below 0. We need therefore to include an offset to take this into consideration. Close the fit window, and re-run a fit, but now with this command:
QSoas> fit-arb im/(1+km/s)+io /with=s:1,exp
Notice the +io bit that corresponds to the addition of an offset current. Load again the base parameters, run the fit again... Your fit window show now look like:
See how the offset current is now much better taken into account. Let's talk a bit more about the parameters:
  • im is \(i_m\), the maximal current, around 120 nA (which matches the magnitude of the original current).
  • io is the offset current, around -3nA.
  • km is the \(K_m\), expressed in the same units as s_1, the first "injected" value of s (we used 50 because the injection is 50 µM CO). That means the value of \(K_m\) determined by this fit is around 9 nM ! (but see below).
  • s_t_1 is the time of the injection of CO (it was in the parameter files you loaded).
  • Finally, s_tau_1 is the time constant of departure of CO, noted \(\tau\) in the equations above.
Taking into account mass-transport limitations
However, the fit is still unsatisfactory: the predicted curve fails to reproduce the curvature at the beginning and at the end of the decrease. This is due to issues linked to mass-transport limitations, which are discussed in details in Merrouch et al, Electrochimica Acta, 2017. In short, what you need to do is to close the fit window again, load the transport.rb Ruby file that contains the definition of the itrpt function, and re-launch the fit window using:
QSoas> ruby-run transport.rb
QSoas> fit-arb itrprt(s,km,nFAm,nFAmu)+io /with=s:1,exp
Load again the parameter file... but this time you'll have to play a bit more with the starting parameters for QSoas to find the right values when you fit. Here are some tips:
  • the curve predicted with the current parameters (use "Update" to update the display) should "look like" the data;
  • apart from io, no parameter should be negative;
  • there may be hints about the correct values in the papers...
A successful fit should look like this:
Here you are ! I hope you enjoyed analyzing our data, and that it will help you analyze yours ! Feel free to comment and ask for clarifications.

About QSoas

QSoas is a powerful open source data analysis program that focuses on flexibility and powerful fitting capacities. It is released under the GNU General Public License. It is described in Fourmond, Anal. Chem., 2016, 88 (10), pp 5050–5052. Current version is 2.2. You can download its source code or buy precompiled versions for MacOS and Windows there.