Home A Primer on TRL Calibration
Post
Cancel

A Primer on TRL Calibration

When using a vector network analyzer (VNA) to measure a device under test (DUT), calibration is a critical step. While there are many calibration methods out there, I believe the thru-reflect-line (TRL) method [1] is one of the best in microwave metrology. Designing a TRL kit is fairly easy and you can manufacture one yourself with PCBs.

In this post, I’ll walk you through the mathematical development of TRL calibration. I have also included a section that will guide you through the design of your own TRL kit. By the end of this post, you’ll be able to implement the calibration yourself and design your own TRL kit with confidence. Also, in my GitHub repository I have Python scripts and measurements to can help you further: https://github.com/ZiadHatab/trl-calibration

If you are more interested in multiline TRL calibration, then check my other post on this topic.

Error Box Modeling of a Two-port VNA

When measuring a two-port DUT with a VNA, we can describe the overall measurement model in terms of cascading the DUT with unknown network boxes. We refer to such a model as the error box model (see Fig. 1).

Illustration of VNA two-port error box model Fig. 1. Illustration of VNA two-port error box model.

The goal of the calibration is to estimate the error boxes. With the obtained estimates, we can remove their effects by an inverse operation (de-cascading), therefore shifting the measurement plane to the DUT. There are, however, some caveats you need to be aware of when modeling a two-port VNA with an error box model:

  1. We assume that the ports of the VNA are isolated from each other, i.e., leakage between the ports is negligible. This is generally true in most VNAs.

  2. Most VNAs measure S-parameters under the assumption that the termination at the non-driving port is reflectionless. However, this is not possible, and we refer to this reflection as “switch terms”. The word “switch term” is quite misleading, as we are actually dealing with reflection due to termination. For the historical reasons behind this terminology, see [2]. You can also check my post on this topic. In short, if you are working with a four-sampler VNA, which is nowadays most VNAs, you can measure the forward and reverse switch terms by connecting any transmissive device between the ports and measuring the following wave ratios: \begin{equation} \Gamma_\mathrm{f} = \left.\frac{a_2}{b_2}\right|_\text{sourced by port 1}, \qquad \Gamma_\mathrm{r} = \left.\frac{a_1}{b_1}\right|_\text{sourced by port 2}\label{eq:1} \end{equation} The equations to correct the S-parameters can be found in references [2] or on my post.

Now that we’ve cleared up the caveats about the error box model, we can delve into the math. The best way to describe the error box model is by the T-parameters, as shown in Fig. 2.

Block diagram of the error box model of a two-port VNA Fig. 2. Block diagram of the error box model of a two-port VNA.

The conversion between the S-parameters and T-parameters is give by

\begin{equation} \bs{T} = \frac{1}{S_{21}}\begin{bmatrix} S_{12}S_{21}-S_{11}S_{22} & S_{11} \\ -S_{22} & 1\end{bmatrix}, \qquad \bs{S} = \frac{1}{T_{22}}\begin{bmatrix}T_{12} & T_{11}T_{22}-T_{12}T_{21} \\ 1 & -T_{21}\end{bmatrix}\label{eq:2} \end{equation}

With the help of the T-parameters definition, we can describe the error box model in Fig. 2 by

\begin{equation} \bs{M}_\mathrm{dut} = \underbrace{k_a \begin{bmatrix} a_{11} & a_{12} \\ a_{21} & 1\end{bmatrix} }_{\text{left error box}}\bs{T}_\mathrm{dut}\underbrace{\begin{bmatrix} b_{11} & b_{12} \\ b_{21} & 1\end{bmatrix}k_b}_{\text{right error box}}\label{eq:3} \end{equation}

The parameters $\{k_a, a_{11}, a_{21}, a_{12}, k_b, b_{11}, b_{21}, b_{12}\}$ are the unknowns that we want to solve for. You can notice that we don’t really need to solve for $k_a$ and $k_b$ separately, we only need their product, thus we can simplify the model by defining $k=k_ak_b$. Therefore, we only need to solve for 7-unknowns, and this simplifies Eq. \eqref{eq:3} to

\begin{equation} \bs{M}_\mathrm{dut} = \underbrace{k_ak_b}_{k}\underbrace{\begin{bmatrix} a_{11} & a_{12} \\ a_{21} & 1\end{bmatrix}}_{\bs{A}}\bs{T}_\mathrm{dut}\underbrace{\begin{bmatrix} b_{11} & b_{12} \\ b_{21} & 1\end{bmatrix}}_{\bs{B}} = k\bs{A}\bs{T}_\mathrm{dut}\bs{B}\label{eq:4} \end{equation}

Solving the Error Boxes with TRL Standards

Now, that we have the model established, we shift our focus on the details of TRL calibration. As the acronym stands for Thru-Reflect-Line, we are working with three standards. The calibration is done in two steps. The first step we use the line and the thru standard to form an eigenvalue problem to solve for normalized calibration coefficients. Then, in a second step we use the reflect standard and the thru standard again to undo the normalization.

The eigenvalue problem

The start by measuring the thru and line standard, described in their T-parameters:

\begin{equation} \bs{M}_\mathrm{thru} = k\bs{A}\begin{bmatrix}1 & 0 \\ 0 & 1\end{bmatrix}\bs{B}, \qquad \bs{M}_\mathrm{line} = k\bs{A}\begin{bmatrix}e^{-l\gamma} & 0 \\ 0 & e^{l\gamma}\end{bmatrix}\bs{B}\label{eq:5} \end{equation}

where $l$ is the length of the line (referenced to the thru) and $\gamma$ is the propagation constant. Note that the thru standard is same as a line standard but with $l=0$. Next, we multiply the inverses of the thru measurement with the line measurement:

\begin{equation} \bs{M}_\mathrm{line}\bs{M}^{-1}_\mathrm{thru} = \bs{A}\begin{bmatrix}e^{-l\gamma} & 0 \\ 0 & e^{l\gamma}\end{bmatrix}\bs{A}^{-1}\label{eq:6} \end{equation}

Eq. \eqref{eq:6} is the eigenvalue problem we are looking for, where $e^{-l\gamma}$ and $e^{l\gamma}$ are the eigenvalues and $\bs{A}$ holds the two eigenvectors. We can also write Eq. \eqref{eq:6} in the following way:

\begin{equation} \bs{M}_\mathrm{line}\bs{M}^{-1}_\mathrm{thru}\begin{bmatrix} a_{11} \\ a_{21}\end{bmatrix} = e^{-l\gamma}\begin{bmatrix} a_{11} \\ a_{21}\end{bmatrix};\qquad \bs{M}_\mathrm{line}\bs{M}^{-1}_\mathrm{thru}\begin{bmatrix} a_{12} \\ 1\end{bmatrix} = e^{l\gamma}\begin{bmatrix} a_{12}\\ 1\end{bmatrix}\label{eq:7} \end{equation}

This means, if apply the eigendecomposition of $\bs{M}_\mathrm{line}\bs{M}^{-1}_\mathrm{thru}$ we should recover the matrix $\bs{A}$ directly, right?! Well, not completely, we have two issues to deal with:

  1. When we apply the eigendecomposition, the order of the eigenvalues is arbitrary. From our point of view, we only get the two eigenvalues as numbers and we don’t know which one represents $e^{-l\gamma}$ or $e^{l\gamma}$.

  2. You should know from your linear algebra 101, that the eigenvectors are unique up to a scalar multiple. That means, we can only solve the coefficients in $\bs{A}$ as normalized coefficients. Think about it this way, look at Eq. \eqref{eq:7}, if we multiply the eigenvectors with a non-zero scalar, it will cancel out automatically, since the scalar exists also on the other side of the equation. That is why any scalar multiple of the eigenvector is still a valid eigenvector.

To resolve the issue with sorting the eigenvalues, we compare the eigenvalues with an estimate we know. That is, we already have an estimate of $\gamma$ to build $e^{-l\gamma}$ and $e^{l\gamma}$. This estimate doesn’t need to be extremely accurate, but should be close enough to the right answer to allow us to distinguish the eigenvalues. Generally, the way this works is that we start with an estimate of $\gamma$ at the first frequency point and solve the eigenvalue problem, and then extract the new $\gamma$ from the eigenvalues. This new $\gamma$ is your new estimate to use in the next frequency point.

Now, the way we solve for the coefficients is by normalizing them such that one element of the vector equals one. The equations in Eq. \eqref{eq:7} become:

\begin{equation} \bs{M}_\mathrm{line}\bs{M}^{-1}_\mathrm{thru}\begin{bmatrix} 1 \\ a_{21}/a_{11}\end{bmatrix} = e^{-l\gamma}\begin{bmatrix} 1 \\ a_{21}/a_{11}\end{bmatrix};\qquad \bs{M}_\mathrm{line}\bs{M}^{-1}_\mathrm{thru}\begin{bmatrix} a_{12} \\ 1\end{bmatrix} = e^{l\gamma}\begin{bmatrix} a_{12}\\ 1 \end{bmatrix} \label{eq:8} \end{equation}

It’s worth noting that $a_{12}$ doesn’t require normalization since the eigenvector is already normalized. You may be wondering why I chose to divide the first eigenvector by $a_{11}$ instead of $a_{21}$, which would also be a valid eigenvector. While this is generally true, $a_{21}$ can be zero, whereas $a_{11}$ cannot be zero as transmission is required for a two-port VNA to function.

Assuming $\bs{v}_1$ is the computed eigenvector corresponding to $e^{-l\gamma}$ and $\bs{v}_2$ is the eigenvector corresponding to $e^{l\gamma}$, we can compute $a_{21}/a_{11}$ and $a_{12}$ as follows:

\begin{equation} a_{21}/a_{11} = \bs{v}_1[1]/\bs{v}_1[0]; \qquad a_{12} = \bs{v}_2[0]/\bs{v}_2[1]\label{eq:9} \end{equation} where the indexing of the vector starts at 0.

Before we continue discussing about how to denormalize the coefficient $a_{21}/a_{11}$, we first need to solve for the right error box $\bs{B}$. Similar to Eq. \eqref{eq:6}, we can also write an eigenvalue problem for $\bs{B}$ as follows:

\begin{equation} (\bs{M}^{-1}_\mathrm{thru}\bs{M}_\mathrm{line})^T = \bs{B}^T\begin{bmatrix} e^{-l\gamma} & 0 \\ 0 & e^{l\gamma}\end{bmatrix}(\bs{B}^T)^{-1}\label{eq:10} \end{equation}

The transpose $()^T$ is used here to push the inverse matrix to the right, otherwise, we would have it on the left side. Now, similar to Eq. \eqref{eq:8}, we can write the eigenvalue problem as:

\begin{equation} (\bs{M}^{-1}_\mathrm{thru}\bs{M}_\mathrm{line})^T\begin{bmatrix} 1 \\ b_{12}/b_{11}\end{bmatrix} = e^{-l\gamma}\begin{bmatrix} 1 \\ b_{12}/b_{11}\end{bmatrix};\qquad (\bs{M}^{-1}_\mathrm{thru}\bs{M}_\mathrm{line})^T\begin{bmatrix} b_{21} \\ 1\end{bmatrix} = e^{l\gamma}\begin{bmatrix} b_{21} \\ 1\end{bmatrix}\label{eq:11} \end{equation}

where \(\bs{B}^T = \begin{bmatrix}b_{11}&b_{21}\\b_{12}&1\end{bmatrix}\).

Now, suppose $\bs{u}_1$ and $\bs{u}_2$ are the computed eigenvectors of $(\bs{M}^{-1}_\mathrm{thru}\bs{M}_\mathrm{line})^T$ that correspond to the eigenvalues $e^{-l\gamma}$ and $e^{l\gamma}$, respectively. Then $b_{12}/b_{11}$ and $b_{21}$ are solved as:

\begin{equation} b_{12}/b_{11} = \bs{u}_1[1]/\bs{u}_1[0];\qquad b_{21} = \bs{u}_2[0]/\bs{u}_2[1]\label{eq:12} \end{equation}

Error terms denormalization

To finalize the calibration, we need to denormalize $a_{21}/a_{11}$ and $b_{12}/b_{11}$, and solve for the 7th error term $k$ (I haven’t forgotten about it). The first what we can do is to use the Thru standard measurement and express the error boxes in terms of their normalized coefficients. This looks something like:

\begin{equation} \bs{M}_\mathrm{thru} = k\underbrace{\begin{bmatrix} 1 & a_{12} \\ a_{21}/a_{11} & 1\end{bmatrix}\begin{bmatrix} a_{11} & 0 \\ 0 & 1\end{bmatrix}}_{\bs{A}}\underbrace{\begin{bmatrix} b_{11} & 0 \\ 0 & 1\end{bmatrix}\begin{bmatrix} 1 & b_{12}/b_{11} \\ b_{21} & 1\end{bmatrix}}_{\bs{B}}\label{eq:13} \end{equation}

By taking the inverse of the normalized matrices, we get:

\begin{equation} \begin{bmatrix} 1 & a_{12} \\ a_{21}/a_{11} & 1\end{bmatrix}^{-1}\bs{M}_\mathrm{thru}\begin{bmatrix} 1 & b_{12}/b_{11} \\ b_{21} & 1\end{bmatrix}^{-1} = \begin{bmatrix} ka_{11}b_{11} & 0 \\ 0 & k\end{bmatrix}\label{eq:14} \end{equation}

So, from Eq. \eqref{eq:14} we are able to solve for $k$. And with $k$ we can also solve for $a_{11}b_{11}$ as

\begin{equation} a_{11}b_{11} = ka_{11}b_{11}/k\label{eq:15} \end{equation}

The next step is to measure the reflect standard. The main property of the reflect standards is that it must be symmetric, that is, both ports must see the same reflect standard. This is illustrated in Fig. 3 below.

error box model of a two-port VNA when measuring a symmetric reflect standard. Fig. 3.: error box model of a two-port VNA when measuring a symmetric reflect standard.

In the figure above, $\Gamma$ is the reflection coefficient of the reflect standard. We don’t know its value, but it should be the same on both ports. In general, it is better to use a short standard, as it radiates less than an open standard. However, any reflective load can be used. The measured reflection seen in the left error box is given by

\begin{equation} \Gamma_a = \frac{a_{12}+a_{11}\Gamma}{1+a_{21}\Gamma}\quad\Longrightarrow\quad \Gamma=\frac{\Gamma_a-a_{12}}{a_{11}-a_{21}\Gamma_a}\label{eq:16} \end{equation}

A similar equation can be obtained for the right error box:

\begin{equation} \Gamma_b = \frac{b_{11}\Gamma-b_{21}}{1-b_{12}\Gamma}\quad\Longrightarrow\quad \Gamma=\frac{\Gamma_b+b_{21}}{b_{11}+b_{12}\Gamma_b}\label{eq:17} \end{equation}

You might be wondering how I came with above equations, and why the measured reflection coefficient in the right error box differs from the left error box? Well, this comes from the famous formula for the input reflection coefficient measured from left-to-right (like with the left error box) \[ \Gamma_{in} = S_{11} + \frac{S_{12}S_{21}\Gamma}{1-S_{22}\Gamma} = \frac{S_{11}+(S_{12}S_{21}-S_{11}S_{22})\Gamma}{1-S_{22}\Gamma} \] The above equation so ubiquities, you can also find it on Wikipedia.
Now, you simply substitute the S-parameters with their corresponding T-parameters. Since above equations is defined from left-to-right, you can’t use it as it is for the right error box (right-to-left). For that, you first flip the S-parameters, i.e., $S_{11} \Leftrightarrow S_{22}$ and $S_{21} \Leftrightarrow S_{12}$, and then substitute with their corresponding T-parameters. That is why, the formula for the left and right reflection coefficient are different.

Since both reflect standards are the same at both ports, we can define the following ratio:

\begin{equation} \frac{a_{11}\Gamma}{b_{11}\Gamma} = \frac{a_{11}}{b_{11}} = \frac{\Gamma_a-a_{12}}{1-(a_{21}/a_{11})\Gamma_a}\frac{1+(b_{12}/b_{11})\Gamma_b}{\Gamma_b+b_{21}}\label{eq:18} \end{equation}

So, now, we have $a_{11}b_{11}$ from the thru measurement and $a_{11}/b_{11}$ from the reflect measurement. Thus, we can solve for $a_{11}$ and $b_{11}$ as follows:

\begin{equation} a_{11} = \pm\sqrt{\frac{a_{11}}{b_{11}}a_{11}b_{11}}; \qquad b_{11} = \frac{a_{11}b_{11}}{a_{11}}\label{eq:19} \end{equation}

To resolve the sign ambiguity, we choose the answer that will give a $\Gamma$ closest to the expected value. For example, if the reflect standard is a short, we expect $\Gamma$ to be close to -1. Of course, you want to take any offset into account if your reflect standard is offsetted by some length.

Now, that we have $a_{11}$ and $b_{11}$, the matrices $\bs{A}$ and $\bs{B}$ are obtained as:

\begin{equation} \bs{A}=\begin{bmatrix} 1 & a_{12} \\ a_{21}/a_{11} & 1\end{bmatrix}\begin{bmatrix} a_{11} & 0 \\ 0 & 1\end{bmatrix}; \qquad \bs{B} = \begin{bmatrix} b_{11} & 0 \\ 0 & 1\end{bmatrix}\begin{bmatrix} 1 & b_{12}/b_{11} \\ b_{21} & 1\end{bmatrix}\label{eq:20} \end{equation}

Everything we derived here is based on the definition of T-parameters as given in Eq. \eqref{eq:2}. Be carful with this, because there are authors that use a different definition for the T-parameters. For example, Michael Steer, in his open access books, he defines the T-parameters as: \[ \bs{T} = \frac{1}{S_{21}}\begin{bmatrix} 1 & -S_{22} \\ S_{11} & S_{12}S_{21}-S_{11}S_{22}\end{bmatrix}; \qquad \bs{S} = \frac{1}{T_{11}}\begin{bmatrix}T_{21} & T_{11}T_{22}-T_{12}T_{21}\\ 1 & -T_{12}\end{bmatrix} \] As you can see, the order in Michael Steer definition is different, but he is consistent with his ordering in the backward conversion. His equations are correct, but if you use it, you want to modify the equations I presented accordingly.
Also, check out Michael Steer’s free books on microwave engineering: https://repository.lib.ncsu.edu/handle/1840.20/36776

Applying the Calibration

Your DUT can be calibrated in terms of T-parameters using the following equation

\begin{equation} \bs{T}_\mathrm{dut} = \frac{1}{k}\bs{A}^{-1}\bs{M}_\mathrm{dut}\bs{B}^{-1}\label{eq:21} \end{equation}

Note that you must first correct for switch terms before applying the calibration in T-parameters. Finally, you can convert back to S-parameters by using eq. \eqref{eq:2}.

On the other hand, it is often desired to convert the obtained error terms to the conventional 12-term VNA model, where the switch term is already included. I won’t derive the equations here, but you can find the details in Joel Dunsmore’s book [3]. The 12 error terms are computed as follows:

  • Single port 3 terms forward direction: \[ E_\mathrm{DF} = a_{12}; \qquad E_\mathrm{SF} = -a_{21}; \qquad E_\mathrm{RF} = a_{11} - a_{12}a_{21} \]

  • Single port 3 terms reverse direction: \[ E_\mathrm{DR} = -b_{21}; \qquad E_\mathrm{SR} = b_{12}; \qquad E_\mathrm{RR} = b_{11} - b_{12}b_{21} \]

  • Remaining 3 forward terms: \[ E_\mathrm{LF} = E_\mathrm{SR} + \frac{E_\mathrm{RR}\Gamma_\mathrm{f}}{1-E_\mathrm{DR}\Gamma_\mathrm{f}}; \qquad E_\mathrm{TF} = \frac{1}{k(1-E_\mathrm{DR}\Gamma_\mathrm{f})}; \qquad E_\mathrm{XF} = 0 \]

  • Remaining 3 terms in the opposite direction: \[ E_\mathrm{LR} = E_\mathrm{SF} + \frac{E_\mathrm{RF}\Gamma_\mathrm{r}}{1-E_\mathrm{DF}\Gamma_\mathrm{r}}; \qquad E_\mathrm{TR} = \frac{kE_\mathrm{RR}E_\mathrm{RF}}{1-E_\mathrm{DF}\Gamma_\mathrm{r}}; \qquad E_\mathrm{XR} = 0 \]

The equations for applying the 12 error terms to obtain the calibrated measurement can be found in [3].

Auxiliary Operations

Below are some operations that are not strictly needed for the calibration itself, but can be useful to extend the calibration beyond the discussion above.

Extracting the propagation constant

After you have computed the eigenvalues and have been able to distinguish between the signs of the exponents. You can calculate the propagation constant $\gamma$ the other way around. Suppose the first eigenvalue $\lambda_1$ is associated with $e^{-\gamma l}$ and the second eigenvalue $\lambda_2$ is associated with $e^{+\gamma l}$. Then we define a new ratio as follows: \begin{equation} e^{2\gamma l} = \lambda_2/\lambda_1\label{eq:22} \end{equation}

We extract the exponent using the logarithmic $e$, which gives us \begin{equation} \log(\lambda_2/\lambda_1) = 2\gamma^{\mathrm{unwrap}} l+ jn2\pi, \qquad \text{for any } n\in\{0,\pm1,\pm2,\ldots\}\label{eq:23} \end{equation}

The extra phase-wrapping factor you see is the result of taking the logarithm of complex numbers, as is unique only in the principle brach $[-\pi,\pi]$. You can read more about the complex logarithm on Wikipedia. For now, we just need to determine the wrap factor to accurately determine $\gamma$. We do this by using the approximation we used to distinguish the eigenvalues. So, the equation to determine $n$ is as follows \begin{equation} n = \mathrm{round}\left( \frac{\Im{\log(\lambda_2/\lambda_1) - 2\gamma^{\mathrm{est}}l}}{2\pi} \right) \label{eq:24} \end{equation}

So the unwrapped $\gamma$ is solved as \begin{equation} \gamma^{\mathrm{unwrap}} = \frac{\log(\lambda_2/\lambda_1) - jn2\pi }{2l}\label{eq:25} \end{equation}

If you are wondering which eigenvalues to take, either from the forward direction (left error box $\bs{A}$), or from the backward direction (right error box $\bs{B}$). The answer is that it doesn’t matter. Both versions give the same eigenvalues, only the eigenvectors change.

Shifting the reference plane

By default, after calibration, the reference plane is set to the center of the thru standard. Therefore, if you were to apply the calibration, it would be referenced to the center of the thru line. However, one of the advantages of the TRL calibration is that you have access to the propagation constant as derived above. With the propagation constant, you can adjust the error boxes to a new reference plane. So we can update the reference plane with the following equations:

\begin{equation} \bs{A}_\mathrm{new} = \bs{A}\begin{bmatrix}e^{-2\gamma d} & 0 \\ 0 & 1\end{bmatrix}; \qquad \bs{B}_\mathrm{new} = \begin{bmatrix}e^{-2\gamma d} & 0 \\ 0 & 1\end{bmatrix}\bs{B}; \qquad k_\mathrm{new} = e^{2\gamma d}k \label{eq:26} \end{equation}

where $d$ is the offset as seen from the port. A positive offset shifts the plane away from the ports, while a negative offset shifts the plane toward the ports.

Illustration of negative plane shift of a TRL calibration done with microstrip line. Fig. 4.: Illustration of negative plane shift of a TRL calibration done with a thru standard as microstrip line.

I will leave it to you as a homework to figure out how I derived these equations. Hint: start by placing a line of length $2d$ between the error boxes $\bs{A}$ and $\bs{B}$.

Renormalize reference impedance

By default, the reference impedance after calibration is set to the characteristic impedance of the line standard. For many types of transmission lines, the characteristic impedance is frequency dependent. The frequency dependency can complicate the interpretation of the calibrated measurements and make it impossible to compare different TRL calibrations of different transmission line types. Therefore, it is often desirable to change the impedance to a different value, most commonly to a frequency-independent value (e.g., 50 ohms).

If the value of the characteristic impedance of the line standard is known, the updated error terms for a new reference impedance can be obtained from the following equations:

\begin{equation} \bs{A}_\mathrm{new} = \frac{\widehat{\bs{A}}}{\widehat{\bs{A}}[1,1]}, \qquad \text{where, } \widehat{\bs{A}} = \bs{A}\bs{Q}^{nm} \label{eq:27} \end{equation}

\begin{equation} \bs{B}_\mathrm{new} = \frac{\widehat{\bs{B}}}{\widehat{\bs{B}}[1,1]}, \qquad \text{where, } \widehat{\bs{B}} = \bs{Q}^{mn}\bs{B} \label{eq:28} \end{equation}

\begin{equation} k_\mathrm{new} = k\widehat{\bs{A}}[1,1]\widehat{\bs{B}}[1,1]\label{eq:29} \end{equation}

The matrix $\bs{Q}^{nm}$ is the impedance transformation matrix [4], which transform the original reference impedance $Z_n$ to the new impedance $Z_m$. The expression for $\bs{Q}^{nm}$ is given as follows (refer to [4] for derivation):
\begin{equation} \bs{Q}^{nm} = \left(\bs{Q}^{mn}\right)^{-1} = \frac{1}{\sqrt{1-\Gamma_{nm}^2}}\begin{bmatrix} 1 & \Gamma_{nm} \\ \Gamma_{nm} & 1\end{bmatrix}, \qquad \text{where, } \Gamma_{nm} = \frac{Z_m-Z_n}{Z_m+Z_n}\label{eq:30} \end{equation}

To learn more about the origin of the transformation matrix, see section “3.7 Change of Reference Impedance” in [4].

In the equations \eqref{eq:27} and \eqref{eq:28} I normalized the new transformed error boxes by their last element. This is because I defined $\bs{A}$ and $\bs{B}$ so that their last element is normalized to one (go back to equation \eqref{eq:3}). As a result, the factor $k$ had to be updated accordingly by the factored out elements.

Design TRL Kit

So this is the most important part of the discussion. Even if you master the math above, if you still make crappy TRL kit, there is no math to fix it. So, the goal of this section is to make you aware of the limitations of TRL kit, which aspects you need to pay attention to, and which of those you need to relax your thinking about.

Line length from frequency range and vice versa

Before proceeding with the discussion below, the term “line length” refers to the difference in length between the line and the thru standard, i.e., the length of the line relative to the thru. By definition, the thru standard has a length of zero. Please keep this in mind when working with a TRL calibration.

Back to the topic at hand. A very important question to ask when designing a TRL kit is: What is the frequency range in which you want the kit to operate?. But an even more fundamental question you should be asking is, Why is there a frequency limit in the first place?

The answer is quite simple. Recall the eigenvalue problem in Eq. \eqref{eq:6}. The eigenvalues depend on $\gamma$, but $\gamma$ itself depends on frequency. This is generally given by the following relation: \begin{equation} \gamma = \alpha + j\beta = \frac{2\pi f}{c_0}\sqrt{-\epsilon_\mathrm{r,eff}}, \label{eq:31} \end{equation} where $\epsilon_\mathrm{r,eff}$ is the relative effective permittivity seen by the wave, which is also a complex-valued number, often written as below (some people use the positive sign convention). \begin{equation} \epsilon_\mathrm{r,eff} = \epsilon_\mathrm{r,eff}^\prime - j\epsilon_\mathrm{r,eff}^{\prime\prime}. \label{eq:32} \end{equation}

The real part of $\gamma$ describes the losses experienced by the wave, while the imaginary part describes the phase of the wave. The frequency limit of TRL comes from the latter term. Therefore, it is easier to explain if we consider the lossless case: \begin{equation} \gamma_\mathrm{lossless} = j\beta = \frac{2\pi f}{c_0}j\sqrt{\epsilon_\mathrm{r,eff}^\prime}. \label{eq:33} \end{equation}

Since $\gamma$ is frequency dependent, the complex exponential terms, i.e., the eigenvalues, will rotate with the frequency in the complex plane. Since the two eigenvalues have different exponent signs, their rotation will be opposite to each other. There will be certain frequencies where the two eigenvalues meet and are equal. This is when the exponent $\gamma l = 0$ or $\gamma l = \pi$. Many people refer to the term $\gamma l$ as the electrical length of the transmission line. A visualization of the eigenvalues rotating in the complex plane with respect to frequency is shown in the figure below.

Illustration of the rotation of the eigenvalues with frequency. Fig. 5.: Illustration of the rotation of eigenvalues in the complex plane with respect to frequency.

I made the animation of the eigenvalues with desmos. What do you think happens to the eigenvalues when we include losses? Funnily enough, when you have high losses, the design criteria for length and frequency get relaxed. Basically, the lossless case is the worst case you can have.

When the eigenvalues are equal, the eigendecomposition collapses and the error boxes vanish into the identity matrix. Consider Eq. \eqref{eq:6}. If we set $e^{-\gamma l} = e^{\gamma l} = \pm 1$, then we get the following equation: \begin{equation} \bs{A}\begin{bmatrix}e^{-l\gamma} & 0 \\ 0 & e^{l\gamma}\end{bmatrix}\bs{A}^{-1} = \bs{A}\begin{bmatrix}\pm 1 & 0 \\ 0 & \pm 1\end{bmatrix}\bs{A}^{-1} = \pm 1\bs{I}_{2\times 2} \label{eq:6mod} \end{equation} Hence, we lose the information on the error box and we cannot recover it. So, clearly, we don’t want our TRL kit to operate in a frequency range where it might cross these critical points. Given the rotational nature of the eigenvalues, we can also visualize the multiband nature of TRL with the sketch below, where we plot the difference between the eigenvalues (hence the sinusoidal function).

We specify a critical phase region, which we call the phase margin. Basically, we want to stay out of this phase margin region. You can see in the sinusoidal plot that the location of the phase margin region is repeated, which also results in multiple passband regions. The choice of what constitutes a good phase margin is up for debate, but the rule of thumb is to use a TRL kit with a phase margin greater than $20^{\circ}$ (the same for both positive and negative sides).

Illustration of phase margin and multi-band nature of TRL calibration. Fig. 6.: Illustration of phase margin and multi-band nature of TRL calibration.

Now, to determine the appropriate length given a frequency band, we write Eq. \eqref{eq:33} as an inequality and multiply it with the length to get the electrical length. We bound the electrical length to the phase margin, and include $\pi$ periodicity to include all bands.

\begin{equation} \pi n + \pi\frac{\varphi}{180}\leq l\frac{2\pi f}{c_0}\sqrt{\epsilon_\mathrm{r,eff}^\prime} \leq \pi n + \left( 1-\frac{\varphi}{180}\right)\pi, \qquad n=0,1,2,\ldots \label{eq:34} \end{equation} where $\varphi$ is the phase margin in degrees. Note that we didn’t define $n$ to be negative because we are limited to positive frequencies.

We can simplify the above equation by canceling $\pi$ on all sides, which gives us \begin{equation} n + \frac{\varphi}{180}\leq l\frac{2f}{c_0}\sqrt{\epsilon_\mathrm{r,eff}^\prime} \leq n + 1-\frac{\varphi}{180} \label{eq:35} \end{equation}

Now we can isolate the frequency term in the middle by multiplying the equation by $c_0/\left(2l\sqrt{\epsilon_\mathrm{r,eff}^\prime}\right)$. \begin{equation} \frac{n + \frac{\varphi}{180}}{2l\sqrt{\epsilon_\mathrm{r,eff}^\prime}}c_0 \leq f \leq \frac{n + 1-\frac{\varphi}{180}}{2l\sqrt{\epsilon_\mathrm{r,eff}^\prime}}c_0 \label{eq:36} \end{equation}

Therefore, given a length $l$, a relative effective permittivity $\epsilon_\mathrm{r,eff}^\prime$, and a phase margin $\varphi$, we can define the lower and upper frequency bounds as follows: \begin{equation} f_\mathrm{min} = \frac{n + \frac{\varphi}{180}}{2l\sqrt{\epsilon_\mathrm{r,eff}^\prime}}c_0, \qquad f_\mathrm{max} = \frac{n + 1-\frac{\varphi}{180}}{2l\sqrt{\epsilon_\mathrm{r,eff}^\prime}}c_0 \label{eq:37} \end{equation} The value of $n$ allows you to move between bands.

Of course, our goal is not only to find the frequency range given an already designed TRL kit, but we also want to design the length given a frequency range specification. We can reformulate Eq. \eqref{eq:37} in terms of $l$ for both frequency bounds: \begin{equation} l = \frac{c_0}{2f_\mathrm{min}\sqrt{\epsilon_\mathrm{r,eff}^\prime}}\left(n + \frac{\varphi}{180}\right), \qquad l = \frac{c_0}{2f_\mathrm{max}\sqrt{\epsilon_\mathrm{r,eff}^\prime}}\left(n + 1 - \frac{\varphi}{180}\right) \label{eq:38} \end{equation}

So, clearly, if you specify any arbitrary $f_\mathrm{min}$ and $f_\mathrm{max}$, you are sure to get different length values. But we all know that length is a constant, so how can we fix this? Since $l$ must be the same for both frequency bounds, we can equate the equations and drop the common terms: \begin{equation} \frac{1}{f_\mathrm{min}}\left(n + \frac{\varphi}{180}\right) = \frac{1}{f_\mathrm{max}}\left(n + 1 - \frac{\varphi}{180}\right) \label{eq:39} \end{equation}

Now we solve for the phase wrapping constant $n$ by rewriting Eq. \eqref{eq:39}, \begin{equation} n = \frac{q - (q+1)\varphi/180}{1-q}, \label{eq:40} \end{equation} where $q = f_\mathrm{min}/f_\mathrm{max}$.

At this point we know that $n$ must be an integer, regardless of the frequency limits and phase margin. Basically, what $n$ tells us is the number of bands that support the specified frequency range and phase margin. Therefore, we need to floor $n$ to the nearest integer. After that, we can calculate the actual phase margin achieved. \begin{equation} n = \left\lfloor \frac{q - (q+1)\varphi/180}{1-q} \right\rfloor \quad\Longrightarrow\quad \varphi = 180 \frac{nq-n+q}{q+1} \label{eq:41} \end{equation}

I should note here that there will be cases where you specify a very wide bandwidth, you will get $n=0$, but you will also get a phase margin smaller than the one you specified. In general, for a desired phase margin of $20^{\circ}$ or higher, you would want $f_\mathrm{max} \leq 8f_\mathrm{min}$. In fact, a more general equation for any phase margin between $0^\circ$ and $90^\circ$ is given by \begin{equation} f_\mathrm{max} \leq \frac{\varphi}{180-\varphi}f_\mathrm{min} \label{eq:42} \end{equation}

In case you are wondering how I derived Eq. \eqref{eq:42}. This is straightforward from Eq. \eqref{eq:41}. Just replace the equality of the phase margin with the $\geq$ relation (we want $\varphi$ to be greater than or equal). Then replace $n=0$ and solve for $q$.

Finally, after you have found $n$ for a given phase margin and frequency limit, and computed back the actual phase margin after flooring $n$. Then you can solve for $l$ using any of the equations in \eqref{eq:38}.

I know this has been too long of a discussion to just calculate $l$. So here is a summary of the steps you should follow when designing the length.

  • Specify a desired phase margin. This ranges from $0^\circ$ to $90^{\circ}$. Typically we choose $\varphi=20^{\circ}$.
  • Specify your target frequency range. Test your frequency limits with the inequality in Eq. \eqref{eq:42}.
  • Use Eq. \eqref{eq:41} to calculate $n$ and the actual phase margin $\varphi$. If you get $n>0$, this means that any value of $n$ that is smaller (down to zero) will give you a higher phase margin (which is better than your specification). In most cases you would use $n=0$.
  • Use any of the equations in \eqref{eq:38} to compute the length.
  • Or just ignore the whole discussion here and use my Python script on GitHub.

Determining the physical length of the thru standard

One common question is how to determine the absolute physical length of the thru standard. While we know that calibration is performed with reference to the center of the thru standard, the length of the standard itself can vary. So, how long should it be? Can it be too short or too long?

To determine the physical length of the thru standard, consider the following:

  1. Crosstalk: ensure connectors and ports are not coupled. If the thru standard is too short, any radiation due to imperfections in the connectors/probes couples with each other. For PCBs with coaxial connectors, the total thru length should be more than three times the length of a connector.

  2. Maximum length: this depends on the losses associated with your medium. If you are dealing with a highly lossy material, keep the thru as short as possible without any coupling problems. If you are dealing with a low-loss material, you can make it longer. As a general rule, make it a little longer than the minimum length that causes no problems.

  3. Tips for specific types of transmission lines:

    • Waveguides: implement the through connections directly with direct flange connection or with a line, and shift the plane back using the propagation constant obtained from calibration.
    • PCBs: introduce a taper between your connector and the final transmission line type and smooth out the traces. Avoid sharp corners if possible.

What should the reflect standard be?

Misconceptions exist about implementing the reflect standard. Some say it must be in the same media as line and thru standards, while others suggest adding an offset or using single-mode propagation. These statements are somewhat true, but can cause confusion when overgeneralized.

The reflect standard must meet three strict properties:

  1. It must be reflective (duhhh~).
  2. It must be symmetrical ( same from both ports).
  3. You must know an estimate of the reflection to resolve sign ambiguity during calibration.

There are possible problems you may encounter during implementation:

  • Replicating the reflect on both ports can be tricky. Poorly designed connectors or interfaces that are used as open standards might resonate in the used frequency band. You can design stable open standards, but it may be better to implement your reflect as a short standard if you are unsure.

  • Adding an offset is common practice to avoid the coupling of the reflect standards with the probes/connectors. For most interface types, an offset is not necessary. However, if you are doing on-wafer probing, you may want to add the offset to be on the safe side, depending on the coupling strength between your probe and the substrate on which the standard is implemented. Adding the offset won’t hurt if you are unsure, but it is unnecessary to do so for every TRL kit.

  • The reflect standard doesn’t have to use the same propagation mode as the line and thru standards. A different mode may make it less stable due to resonance or similar effects, but this doesn’t need to be the case. You can use leave the cable open on both ends (if stable) or using a coaxial short standard from your SOLT box. Don’t overthink the propagation mode of your reflect standard.

  • It’s important to know an estimate value of the reflect standard to resolve the sign ambiguity. Most people make it on the same media as the line and thru standards, making it easy to predict its reflection coefficient. One useful trick is to approximate the reflectance at the first frequency point and use the new estimate at each new frequency point. This is helpful when implementing the reflect as open or short standards, which have a reflection coefficient of approximately 1 and -1 respectively, at low frequencies.

References

[1] G. F. Engen and C. A. Hoer, “Thru-Reflect-Line: An Improved Technique for Calibrating the Dual Six-Port Automatic Network Analyzer,” in IEEE Transactions on Microwave Theory and Techniques, vol. 27, no. 12, pp. 987-993, Dec. 1979, doi: 10.1109/TMTT.1979.1129778.

[2] R. B. Marks, “Formulations of the Basic Vector Network Analyzer Error Model including Switch-Terms,” 50th ARFTG Conference Digest, 1997, pp. 115-126, doi: 10.1109/ARFTG.1997.327265.

[3] J. Dunsmore, Handbook of Microwave Component Measurements. John Wiley & Sons, Ltd, 2020, doi: 10.1002/9781119477167

[4] Marks, R. and Williams, D. (1992), A General Waveguide Circuit Theory, Journal of Research (NIST JRES), National Institute of Standards and Technology, Gaithersburg, MD, doi: 10.6028/jres.097.024

This post is licensed under CC BY 4.0 by the author.

-

VNA Switch Terms

Trending Tags