Noise Figure and Signal to Noise Ratio

Noise Sources in Communications and Radar Systems

Intrinsic Sources of Noise

Extrinsic Sources of Noise

Thermal, Nyquist, Johnson or 'kTB' Noise

Conventions, Assumptions and Limitations

Additive White Gaussian Noise

The Derivation AWGN from Black Body Radiation

Operating, Noise and Measurement Bandwidths

Noise Power Spectral Density and Frequency Planning

Noise Factor and Noise Figure

Definition of Noise Factor

Standard Noise Temperature \(T_0 = \) 290 \(K\)

Signal to Noise Ratio in Cascaded Systems

Excess Noise Power and Equivalent Input Noise Temperature

Under Construction

Noise Factor of a Matched Attenuator

AWGN in Cascaded Systems

Friis's Formula for Cascaded Stages

Noise Measure

Antenna Noise Temperature

Measuring Noise Factor Using the Y-Factor Method

Calibration

Measurement

Frequency Conversion

DUT Without Internal Frequency Conversion

DUT With Internal Frequency Conversion

Low Noise Antenna Receiving System

Under Construction

References

****************************************************************************************************

Noise Sources in Communications and Radar Systems

Electrical noise (noise) in communication and radar systems can be divided into two groups: intrinsic (internal to the hardware) and extrinsic (external to the hardware). In each case 'the hardware' refers to the susbsystem under consideration, such as an amplifier, receiver, filter, or a cascade of 2-port devices. Noise affects all electronic equipment but, depending on its application, some types of noise may be insignificant and others may be critical. In new designs a proper assessment of all forms of noise is required.

Some forms of intrinsic and extrinsic noise are described in the following paragraphs [1] [5] [15].

Intrinsic Sources of Noise

Extrinsic Sources of Noise

Thermal, Nyquist, Johnson or 'kTB' Noise

Thermal, Nyquist, Johnson or 'kTB' noise, generically known as additive white Gaussian noise (AWGN) is the type associated with noise factor and noise figure definitions and therefore signal to noise ratio performance. It is generated by all electrical equipment because it is present at temperatures above Absolute Zero. Its mean power is a function of the absolute temperature of the noise source and the effective noise bandwidth of the electrical equipment which generates it [4] [5] [7] [8] [10] [11] [12].

The schematic in Figure 1.3-1 shows how AWGN noise power may be measured for an AWGN source operating at an absolute temperature of \(T\) kelvin (\(K\)).

Figure 1.3-1 The ideal measurement of an AWGN source at an equivalent noise temperature of \(T\) \(K\). The bandpass filter is assumed to be an ideal 'brickwall' type.

All components are assumed to be in thermal equilibrium at temperature \(T \,K \) and well matched. AWGN is observed up to frequencies well beyond 100 \(GHz\) so it is present in all electrical equipment operating within the limits of the current technology [7] [13]. The components are assumed to be operating within this range. The AWGN mean power (\(P_n\)) watts (\(W\)) is the product of Boltzmann's constant (\(k = 1.38 \times 10^{-23} J/K\)), the absolute temperature \(T\) \(K\) and the equivalent noise bandwidth of the bandpass filter, \(B_N\) hertz (\(Hz\)). The bandpass filter represents the practical operating bandwidth of the electrical device under test (DUT). In most cases the DUT will be a common 2-port device carrying signals such as an amplifier, attenuator, filter or mixer. If the frequency range has a bandwidth \(B_N\) \(Hz\), then \(P_n\) is given by (1.3-1).

\[ P_n = k \, T B_N \,\,\, W\]
(1.3-1) The formula for AWGN power \(P_n\) \(W\) generated at an absolute temperature \(T\) \(K\) in a noise bandwidth \(B_N\) \(Hz\).

Conventions, Assumptions and Limitations

There are many types of electrical noise. In this article we are only considering AWGN, arguably the most important noise affecting the sensitive front ends of receiving systems.

Additive White Gaussian Noise

We have made many references to additive white Gaussian noise (AWGN) so some more background information on the physics behind this is included in the following sections.

The Derivation AWGN from Black Body Radiation

For a resistor \(R\) ohms (\(\Omega\)) at an absolute temperature \(T \) kelvin (\(K\)), the random and independent motion of large numbers of electron charges results in voltage fluctuations across the resistor with a Gaussian distribution. The average vector voltage of the AWGN is zero, but the mean square voltage is finite and proportional to the mean power of the AWGN for a given bandwidth. The root mean square (RMS) value of the noise voltage, \(V_n\) volts (\(V\)), is given by Planck's black body radiation law, (3-1) [1] [7] [9].

\[V_n = \sqrt{\frac{4 h f B R}{e^{\frac{h f}{k T}}- 1}} \, \,volts\]
(3-1) The RMS AWGN voltage, derived from Planck's black body radiation law.

where:

The instantaneous noise voltage within a defined noise bandwidth is not coherent or sinusoidal, its resultant vector or phasor voltage being zero, but the mean square voltage is finite and proportional to the mean AWGN power. The adjective 'additive' in the AWGN acronym describes how the noise powers of contiguous frequency bands may be simply added to yield the noise power of the aggregate band. This is a very useful property of AWGN. We will first apply this equation exactly and then simplify it to derive the \(k \,T B\) form.

With an average (vector) value of zero, we cannot extract a phasor value from (3-1). However, the associated noise power is finite and proportional to the mean square voltage. To calculate the AWGN power, we need to incorporate the noisy resistor \(R\) into a perfectly matched Thevenin equivalent constant voltage circuit as shown in Figure 3-1. This includes a noise-free source resistor also of resistance \(R\) and an ideal 'brickwall' bandpass filter (BPF). We will see that the 'white' property of AWGN extends to beyond 100 \(GHz\) so the BPF is required to restrict the noise bandwidth under consideration (\(B_N\)) to some useful range relevant to the 2-port devices in use [7] [13].

Figure 3-1. The Thevenin constant voltage source for the AWGN source connected for maximum power transfer to a load resistance \(R\).

Figure 3-1 shows that, to calculate the mean AWGN power (\(P_n\)) we must use, not \(V_n\) but \(V_n/2\) (3-2):

\[P_n = \frac{V_n^2}{4R} = \frac{h f B_N}{e^{\frac{hf}{kT}}-1} \,\,watts\]
(3-2) The mean AWGN power \(P_n\) watts within the noise bandwidth \(B_N\) hertz at a frequency \(f\) hertz and absolute temperature \(T\) kelvin.

We have also substituted for \(V_n\) from (3-1). Popular units for describing AWGN power spectral density, say \(PSD1\), is decibels relative to one milliwatt per hertz (\(dBm/Hz\)). \(P_n\) (in watts) must initially be multiplied by \(10^3\) to convert to milliwatts, then divided by \(B_N = 1 \, Hz\). Then, ten times the logarithm to base 10 of this value gives the result in these units (3-3).

\[PSD1 = 10 \log_{10} \left[10^3 \left(\frac{P_n}{B_N}\right)\right] \,\,dBm/Hz\]
(3-3) The AWGN power spectral density as a function of absolute frequency at a temperature of \(T = 290 K\).

We are using \(T=T_0\) again for which \(PSD1\) is the red plot in Figure 3-2

Figure 3-2. From (3-2) and (3-3), plots of AWGN power spectral density \(PSD1\) and \(PSD2\) against absolute frequency.

The \(PSD1\) plot shows the AWGN noise power density to be constant at about -174 \(d B m/Hz\) up to about 1000 \(G H z\) after which it tails off steeply. However, the current generation of electronics hardware based on conductors (circuit and waveguide technology) is included within this frequency range and the tail-off should not concern us too much. The flat region of \(PSD1\) demonstrates the 'white' property of AWGN, the analogy being with the optical part of the electromagnetic spectrum and its component colours.

Now we will make a very useful simplification to obtain the less cumbersome \(k T B\) form of equation derived from (3-2). For frequencies below about 100 \(GHz\) and at operating temperatures close to 'room temperature', approximately \(T = 290 \, K\), \(h f \, \lt \lt k T \) is valid and the first 2 terms of a Taylor series gives (3-4) [1]:

\[e^{\frac{hf}{kT}}-1 \approx \frac{hf}{kT} \]
(3-4) The approximation used to estimate the RMS AWGN voltage at realistic frequencies and temperature for most electronic parts and systems.

Using (3-4) in (3-1) results in (3-5), also known as the Raleigh-Jeans approximation [1].

\[V_n = \sqrt{4 k T B R} \, \, volts\]
(3-5) The simplified and practical equation for the RMS voltage of AWGN across the RF and microwave frequency range at typical temperatures.

This equation is sufficiently accurate to use for all electronic equipment at the current limits of the technology. For very low temperatures and/or very high frequencies, using (3-1) directly may be slightly more accurate. The same procedure applied for the 'exact' equation using the Thevenin equivalent circuit shown in Figure 3-1 may then be applied to arrive at the well known AWGN equation (3-6).

\[ P_n = k \, T B_N \, \, watts\]
(3-6) The mean AWGN noise power as a function of noise bandwidth \(B\) and absolute temperature \(T\).

We know that \(k = 1.38 \times 10^{-23} \, J/K\) is Boltzmann's constant. \(T\) is the absolute temperature of the AWGN source. I have added the subscript '\(N\)' to the bandwidth symbol '\(B\)' to emphasise that it refers to the (AWGN) noise bandwidth as opposed to say, the -3 \(dB\) CW (VNA-based) bandwidth, or to some other definition of bandwidth [9] [15]. The noise bandwidth of a real bandpass filter is the bandwidth of the equivalent rectangular ('brickwall') response shape as it affects AWGN. For most cases, the relevant noise bandwidth is that which affects the whole cascade of 2-port devices from the signal input to immediately before the detector. This is examined further in Section 3.2. An important and useful property of this equation is that it is not a function of the absolute frequency, but it is a function of noise bandwidth itself. The AWGN may be expressed as a noise power density (NPD) with respect to frequency or spectral power density (SPD) by dividing (3-6) by \(B_N\) (3-7).

\[ \frac{P_n}{B_N} = k \, T \, \, watts/hertz\]
(3-7) The mean AWGN spectral power density (SPD) or noise power density (NPD).

After again converting this to units of \(dBm/Hz\), it is included as \(PSD\,2\), the blue plot in Figure 3-2. Comparing the red and blue plots shows that the simpler version of the AWGN equation, (3-6), is adequate for frequencies up to around 1000 \(GHz\). Therefore, a common condition of using (3-6) is that the absolute frequency must be less than about 100 \(GHz\).

The next question is usually, "What noise bandwidth should I actually choose for a particular system?". This decision needs to consider the devices under test and the objectives against the limitations of the test equipment. These will be discussed in Section Section 3.2

Operating, Noise and Measurement Bandwidths

A 2-port device under test (DUT) has both an operating bandwith and an equivalent noise bandwidth, or simply 'noise bandwidth' \(B_N\). We also need to know its various other properties such as input and output operating frequency ranges, its channelisation scheme and modulation characteristics. Noise bandwidth should be used in any analyses which relate to AWGN.

The operating bandwidth of a DUT is usually measured using a continuous wave (CW) frequency source, typically part of a correctly set up and calibrated vector network analyzer (VNA). VNA measurements may be repeated across a range of stepped CW frequencies, usually selected to extend slightly beyond the expected operating bandwidth. The noise bandwidth of the DUT may be measured empirically using a broadband AWGN source connected to the DUT input and a sensitive power meter connected to the output. This is shown in Figure 3.2-1.

Figure 3.2-1 The AWGN noise bandwidth of a device is the bandwidth of a hypothetical rectangular bandpass filter which may be used to replace the DUT to give the same output noise power.

The equivalent noise temperature used for the AWGN source (\(T \, \, K\)) and the power meter sensitivity must be chosen appropriately for the DUT bandwidth. Achieving this is sometimes challenging. Alternatively, the noise bandwidth may be calculated if accurate DUT transmission data are known. Most modern, high specification VNAs enable us to export this electronically into 'number crunching' applications for further processing such as Excel® and Matlab®.

As we mentioned, the noise bandwidth of a DUT is the bandwidth of the equivalent rectangular (or 'brickwall') passband filter, referenced to the same transmission amplitude and center frequency, that will transmit the equivalent mean AWGN power if it replaced the DUT. This may be shown using the normalized example transmission responses in Figure 3.2-2 represented by the blue and red continuous plots. This demonstrates how two different devices may have the same VNA (CW-based, -3 \(dB\)) bandwidths but differing noise bandwidths.

Figure 3.2-2 Two DUT transmission against frequency characteristics (red and blue continuous plots) with the same VNA half power (-3 \(d B\)) bandwidths may have differing shapes and therefore differing AWGN noise bandwidths shown by the respective, rectangular dotted responses.

A noise bandwidth may be calculated for any DUT transmission response but those with near symmetry about a center frequency, like either of those shown in Figure 3.2-2, are typical. In this example, the center frequency is 100 \(MHz\) coincident with the minimum insertion loss. Here, the response is normalized so this corresponds to a magnitude of one. This frequency will be used as a reference for the actual response and the noise bandwidth response.

Referring to Figure 3.2-1, if the effective noise temperature of the AWGN source is \(T = T_0 = 290 K \), the noise power spectral density (NPD) from it, feeding the DUT, is \(k \,T_0 \,\, \,W/Hz \). The total mean noise power passing through the DUT \(P_T\) is given by (3.2-1), where \(G(f)\) is the function describing the linear scalar transmission of the DUT with respect to frequency. This is equivalent to the area under the response plot.

\[ P_T = \int_0^\infty k \, T_0 G(f) \, df = \int_{0 \, Hz}^{100 \, GHz} k \, T_0 G(f) df \,\,\, watts \]
(3.2-1) The mean AWGN power transmission through the actual DUT.

A practical upper integration limit of 100 \(GHz\) was chosen based on the Rayleigh-Jeans criteria from Planck's black body radiation law, giving a relatively flat AWGN NPD. This was discussed in Section 3.1 and allows us to use the simplified AWGN equation (3-6). After substituting the real DUT for the hypothetical noise bandwidth (\(B_N\)) filter, noting that the transmission amplitude is now referenced to its value at the center frequency of the real DUT, \(G(f_0)\), \(P_T\), again the area under the response is (3.2-2).

\[ P_T = k \, T_0 B_N G(f_0)\,\,\, watts \]
(3.2-2) The mean AWGN power transmission through the equivalent noise bandwidth filter.

Normalized responses, say \(G_N(f)\), of the type shown in Figure 3.2-2 are defined as (3.2-3).

\[ G_N(f) = \frac{G(f)}{G(f_0)} \]
(3.2-3) The DUT scalar and normalized response as a function of frequency.

By equating (3.2-1) with (3.2-2), and using the definition in (3.2-3) gives the following expression for \(B_N\) (3.2-4).

\[ B_N = \int_{0 \, Hz}^{100 \, GHz} G_N(f) df \,\,\, hertz \]
(3.2-4) The noise bandwidth \(B_N\) in terms of the normalized transmission amplitude against frequency response \(G_N(f)\), such as with the examples shown in Figure 3.2-2.

The example DUTs shown in Figure 3.2-2 have -3 \(dB\) bandwidths of about 5 \(MHz\) and we suggested that these may have been measured using a VNA. In order to perform AWGN measurements on these to a similar frequency resolution, such as those necessary for noise figure, the noise figure meter (NFM) would need to step a narrow measurement noise bandwidth \(B_N\) transmission response across the operating band in small increments. Suppose the operating bandwidth extends from a lower frequency \(f_L\) to an upper frequency \(f_U\), plus a suitable margin, then we would require (3.2-5).

\[ B_N \lt\lt f_U-f_L \]
(3.2-5) For the best resolution the noise bandwidth \(B_N\) must be much less than the operating bandwidth, \(f_U - f_L \).

So, for an operating bandwidth of around 5 \(MHz\) we might expect the \(B_N\) of the NFM to be perhaps a few tens of kilohertz. If you have the budget, this can be done with some of the latest noise figure analyzer (NFA) test equipment [20]. In fairness to the NFA manufacturers, I would add that the latest NFAs generally have DSP-based, real-time architectures, numerous add-on options such as spectrum analysis and resolution bandwidths down to 1 \(Hz\). Otherwise, we will be using the Hewlett Packard (HP) Noise Figure Meter (NFM) model HP 8970B with the switched AWGN 14 \(dB\) ENR noise source HP 346B [16]. These instruments may be considered 'legacy' but they were widely respected industry standards for many years. The HP 8970B has an input frequency range from 10 \(MHz\) to 1600 \(MHz\) with a bandwidth of 'approximately 4 \(MHz\)'. This approximates to a noise bandwidth also of 4 \(MHz\), with a center frequency tunable across the input frequency range. For now we are assuming the DUT is a device without internal frequency conversion, such as an amplifier, attenuator or filter. Also, the DUT operating frequency range is within the frequency ranges of both the NFM and the noise source (we are not including any frequency conversion external to the DUT). DUT measurements with frequency conversion will be addressed in Section 5.5. Figure 3.2-3 shows how the noise bandwidth \(B_N\) of the NFM may be swept across the operating bandwidth of the DUT \(f_U - f_L\) for a typical noise figure measurement.

Figure 3.2-3 This transmission against frequency response shows that an accurate noise factor measurement with good resolution will be achieved across the DUT operating bandwidth and beyond if the noise measurement bandwidth \(B_N\) is a small fraction of the operating bandwidth \( \left( B_N \lt \lt f_U - f_L \right) \).

Noise Power Spectral Density and Frequency Planning

Noise power spectral density (PSD) or simply noise power density (NPD) are alternative descriptions of the same property, the AWGN power density with respect to (absolute) frequency. As discussed in Section 3.1, as long as the operating frequency is less than about 100 \(GHz\) and temperature stable, a calibrated AWGN source will generate a constant NPD which is not a function of frequency [7] [13]. Most practical receiving systems operating with today's technology will have operating bandwidths which are much less than 100 \(GHz\). The majority of traffic channel bandwidths in communications and radar have evolved to a range from about 1 \(MHz\) to 100 \(MHz\). With careful frequency planning, the operating bandwidth of a DUT such as a low noise amplifier (LNA) can be designed to cover several such traffic channels. The same product can potentially be selected by several customers such as telecommunications service providers, each with slightly different traffic allocations.

The Y-factor method, described in Section 5.4, is often used by NFMs and NFAs for accurate and quick noise figure and noise gain measurements. This also requires a 'noise source' such as the HP 346B. The HP 346B and other similar devices, actually comprises two AWGN sources, each with accurately known noise temperatures and controlled by the associated NFM or NFA. The 'white' adjective in the AWGN acronym, being an analogy with white light in the optical spectrum, means that the AWGN power density across the operating frequency range is nominally constant. Actually, due to practical design limitations there is usually some small variation with frequency but this can be corrected for during calibration.

A popular unit for AWGN NPD is the logarithmic: 'decibels relative to one milliwatt per hertz', with the symbol \(dBm / Hz\). This does not imply that literally measuring the AWGN in a 1 \(Hz\) bandwidth is necessary (although it is possible with some of the latest high specification NFAs) [20]. Substituting the standard noise temperature of \(T = T_0 = 290 K\) and noise bandwidth \(B_N = 1 Hz\) into the AWGN power equation (3-6) tells us that the AWGN power in 1 \(Hz\) is -174 \(dBm\). This is difficult to measure directly, but the white property of AWGN here is useful and this is one example of its use. Widely available and relatively inexpensive noise figure equipment will typically have measurement (noise) bandwidths in the order of a few megahertz. The white property of AWGN means that the noise power measured in, say a 4 \(M H z\) noise bandwidth, may simply be linearly scaled down to narrower bandwidths such as 1 \(H z\) to convert it to the \(dBm / Hz\) unit. This example would be a linear factor of \(0.25 \times 10^{-6}\) or a logarithmic value of \(10\log_{10}(0.25\times 10^{-6}) \, dB\) or \(-66 \, dB\). To obtain such a value accurately would require the larger noise bandwidth to be known precisely.

Noise Factor and Noise Figure

In many of the references there is some inconsistency in the use of the terms noise factor (\(F\)) and noise figure (\(f\)). As described in Section 2 under 'Linear and Logarithmic Units', we will keep with the convention of \(F\) being the linear form, expressed as a unitless positive real number greater than or equal to 1, and \(f\) being the logarithmic form, expressed as a positive value in \(dB\).

Definition of Noise Factor

Noise factor will be described with the help of the schematic showing the 2-port device under test (DUT), which does not have frequency conversion, with the test equipment shown in Figure 4.1-1.

Figure 4.1-1 A simplified schematic to demonstrate the procedure for measuring the noise factor of a 2-port device.

It is assumed that the DUT is well matched at both ports to the test equipment and all devices are in thermal equilibrium. The ambient temperature of the DUT should be controlled and stabilized to the standard noise temperature, \(T_0 = 290 K\). The dotted box contains a sensitive, RMS sampling power meter and a bandpass filter with a tunable center frequency to represent a simplified noise figure meter (NFM). The AWGN source is broadband, selected to cover a frequency range in excess of the operating bandwidth of the DUT. The frequency range of the NFM \(f_M\) must be within the operating frequency range of the AWGN source and cover the operating frequency range of the DUT, \(f_L\) to \(f_U\). The noise bandwidth \(B_N\) for all measurements is defined by the tuneable bandwidth of the NFM. For the best frequency resolution, \(B_N\) must be much less than the DUT operational bandwidth as was described in Section 3.2. Although this configuration does not use a CW source signal similar to a VNA, under the conditions described it may use the AWGN source to measure gain. For calibration purposes, the DUT may be removed and the combined sources connected directly to the NFM including any cables and adaptors necessary for the measurement. For measurements, the DUT is inserted as shown. The calibration and measurement configurations allow both linear gain of the DUT (\(G\)) and its noise factor (\(F\)) to be measured. Under these conditions, \(G\) is equivalent to the linear gain that could be measured using CW with a VNA.

The DUT input CW (signal) power is \(S_i\) and the input AWGN power is \(N_i\). The DUT output CW power is \(S_o\) and the output AWGN power is \(N_o\). All powers are linear, specified in watts (\(W\)).

The noise factor definition for a matched 2-port device like this DUT is the ratio of the signal to noise ratio (SNR) at the input divided by the SNR at the output. The linear gain is \(G = S_o / S_i\) and \(N_i\) equates to \(P_n\) in (3-6). \(F\) is therefore expressed by the following equation (4.1-1).

\[ \ F = \frac{\frac{S_i}{N_i}}{\frac{S_o}{N_o}} = \frac{N_o}{G N_i} = \frac{N_o}{G k T B_N} = \frac{N_o}{G k T_0 B_N} \]
(4.1-1) The definition of noise factor (\(F\)) in terms of the input and output signal and noise powers and the linear gain.

\(B_N\) is not the noise bandwidth of the DUT alone, which would be a value similar to its operating bandwidth, but a much smaller value chosen according to (3.2-5). As \(F\) is a function of \(T\), the latter must be specified and appropriate to the system under consideration. Linear (small signal) conditions are assumed. Therefore the DUT gain has an equivalent effect on both signals and noise. Here we have assumed the ambient temperature of the source and DUT to be the standard noise temperature as is frequently used in datasheets, \(T = T_0 = 290 \, K\). For antenna receive systems or unusually hot or cold environments, the actual noise factor will change accordingly.

For AWGN related parameters (noise factor, gain and SNR), 2-port devices may be represented schematically like the example shown in Figure 4.1-2.

Figure 4.1-2 The AWGN schematic for a matched 2-port device in thermal equilibrium. \(N_x\) is the excess AWGN power of the device referred to the output.

Not to be confused with similar schematics for the Thevenin and Norton equivalent circuits, this AWGN schematic displays the real DUT input and output mean noise powers, and how they relate to each other. AWGN phase is random and therfore unpredictable. Power has a square law function of instantaneous voltage so does not have a phase. Therefore we can only use mean power levels or mean square voltages to describe AWGN levels. These are real powers so they are additive, hence the first letter of the AWGN acronym. Within a cascade, the stages are assumed to be well matched so there are no issues with standing waves or reflected power. In fact there are no 'waves' as nothing is sinusoidal and there is no phase reference as would be provided by a sinusoidal source. That is why we have given the symbol for the AWGN source a hatched appearance.

Standard Noise Temperature \(T_0 = \) 290 \(K\)

From (4.1-1) \(F\) is a function of \(N_i\). With the AWGN source, we know from (3-6) (\(N_i = P_n\)) that \(F\) is also a function of the effective noise temperature of the AWGN source \(T\). This shows an important property of the noise factor definition. The noise factor is actually dependent on the NPD (3-7) from the source, feeding the DUT input. In datasheets, the effective temperature of the AWGN source is often standardised to 290 \(K\). This is an evolution of 'average' room temperature adopted for convenience in the design of low noise amplifiers in the early years of satellite communications and radio astronomy. It is now an Institution of Electrical and Electronics Engineers (IEEE) accredited definition [19]. It assumes that the equivalent input noise temperature (\(T_0\)) is 290 \(K\) which, using the discussion in Section 3.1 and (3-7) gives a NPD of -174 \(dBm/Hz\). There was clearly an assumption that, in most cases, the low noise device under consideration was operating at approximately room temperature.

Take the example of a narrow beam receiving antenna, typical of the type used for satellite communications [27]. This is highly directional and designed to only collect emissions from a small near circular aperture in space when pointing at the position of the satellite. We will assume it is an ideal antenna with zero loss and perfectly matched to the first stage of the receive cascade, a low noise amplifier (LNA). If the satellite transmitter was 'switched off' the ground antenna would then collect sky noise source(s), which approximate to AWGN, from the same direction as the satellite. With the exception of possible periods of Sun interference, the only AWGN sources are from distant stars. Distances from these are huge, usually measured in light-years. Furthermore, the AWGN power flux density (PFD) of noise from a distant star will suffer 'spreading' to an inverse square law towards the Earth [24] [25]. The linear SI unit for PFD is watts per square meter (\(W/m^2\)). The combined PFDs of noise sources like these from many stars arriving at the Earth have been measured accurately. Some such measurements arrived at an equivalent antenna noise temperature sometimes below 50 \(K\) [23]. If we had tested the LNA designed for this antenna in the laboratory, assuming the standard noise temperature of \(T = T_0 = 290 \, K\), the SNR after implementation would actually be greater. Therefore, when designing low noise receiving systems like these, we have to make the necessary corrections. In Section 6 we will look at a similar example.

Antennas used for terrestrial communications tend to have noise temperatures closer to 290 \(K\) because the beam shapes typically include substantial natural ground sources which are much closer to this temperatue than, say 50 \(K\). Also, they are physically much closer, so the beam spreading due to the inverse square law is less significant. However, assuming that noise from the Sun is not within any part of the beam, this is offset by the upper part of the beam which will usually see sources at colder temperatures.

Signal to Noise Ratio in Cascaded Systems

Excess Noise Power and Equivalent Input Noise Temperature

Figure 4.1-2 shows a schematic for a 2-port 'noisy' DUT together with symbols for the parameters which were described in Section 4.1. The adjective 'noisy' simply means that the DUT is a real, imperfect device which adds some finite level of AWGN to the signal applied to the input. The AWGN will only be zero in a device that is at zero degrees absolute (\(0 \, K\)), so all real devices are 'noisy'.

The first task is to choose a measurement or noise bandwidth \(B_N\). We will choose 4 \(MHz\), as we have the HP 8970B NFM in mind as an example. Ideally this should be much less than the operating bandwidth of the DUT but we can live with it being just slightly less provided we are very careful with the choice of the measurement frequency (the center frequency of the tuneable bandpass filter device within the NFM). We need to avoid inadvertantly tuning to the side of the passband response. Examples of modern, high specification noise figure analyzers (NFAs) can have measurement bandwidths as small as 1 \(Hz\) [20]. The 'additive' and 'white' properties of AWGN mean that the AWGN measured in a wide bandwidth like 4 \(MHz\) can be scaled linearly down to a narrower bandwidth. However, this conversion must be performed with caution as the measurement bandwidth transmission characteristic needs to be measured very accurately.

We must also consider the smallest service channel bandwidth carried by the DUT which requires testing. For example, a DUT designed with an operating bandwidth of 500 \(MHz\) targetted at several 4G and 5G digital cellular service providers may be required to carry channel bandwidths as small as 5 \(MHz\). High resolution noise figure (and gain) data would be useful for potential customers to consider.

Using (3-6) with \(N_i = P_n\) and \(T_0 = T\), gives us the noise factor definition (4.4-1).

\[ \ F = \frac{N_o}{G N_i} = \frac{N_o}{G k \, T B_N} = \frac{N_o}{G k \, T_0 B_N} \]
(4.4-1) The noise factor definition (4.1-1) applies to a stabilized ambient temperature \(T_0\) of 290 \(K\).

We are assuming the standard temperature of \(T = T_0 = 290 \, K\) is used for the AWGN source, so essentially a 'room temperature' application. In this noisy DUT the output noise power comprises 2 additive components: the amplified input noise \(G k T_0 B_N\) and the noise added by the DUT itself or 'excess' AWGN, say \(N_x\), referred to the output (4.4-2). \(N_x\) does not depend on either the input signal or noise power, it is only a function of the DUT electronics hardware and the temperature at which it is operating, in this case \(T_0\).

\[ N_o = G k T_0 B_N + N_x \]
(4.4-2) The output noise power comprises the amplified input noise power \(GkT_0B_N\) and the excess noise power \(N_x\), referred to the output.

Substituting (4.4-2) into (4.4-1) gives (4.4-3)

\[ F = \frac{N_o}{G N_i} = \frac{N_o}{G k \, T_0 B_N} = \frac{G k \, T_0 B_N + N_x}{G k \, T_0 B_N} = 1 +\frac{N_x}{G k \, T_0 B_N} \]
(4.4-3) The noise factor definition, including the contribution of the excess noise power \(N_x\), referred to the output.

As \(N_x\) is an AWGN (output) power, it may be represented as an equivalent temperature using (3-6) in the same noise bandwidth \(B_N\). It is most commonly represented as an equivalent input noise temperature \(T_e\) but could equally be represented as an equivalent output noise temperature. We will chose the former so that the \(GkB_N\) parts cancel out. Therefore, to express \(N_x\) in terms of \(T_e\) we need (4.4-4).

\[ N_x = G k T_e B_N \]
(4.4-4) The output excess noise power \(N_x\) in terms of the equivalent input noise temperature \(T_e\).

Substituting \(N_x\) from (4.4-4) into (4.4-3) gives us the very important noise factor relationship (4.4-5).

\[ F = 1 + \frac{N_x}{GkT_0B_N}= 1 + \frac{GkT_e B_N}{GkT_0B_N} = 1+\frac{T_e}{T_0} \]
(4.4-5) The noise factor in terms of the equivalent input noise temperature \(T_e\) and the standard noise temperature \(T_0\).

Under Construction

Noise Factor of a Matched Attenuator

An attenuator, unlike an amplifer, is not connected to a power supply and therefore the source of AWGN is limited. This might sound counter-intuitive because we know that an attenuator comprises a network of accurately trimmed resistors designed to attenuate a signal (or noise) whilst maintaining a good input and output match. Yes, resistors generate noise if they are acting as a source but they also attenuate noise in the same way that they attenuate signals. An attenuator is a 2-port device just like an amplifier and does not behave like a one port AWGN source, for example a noise source or an antenna. To calculate its noise factor, we simply apply the same definitions used for any other 2-port linear device. In fact it is more straightforward than an amplifier: because it is passive, we can call upon conservation of power.

Figure 4.6-1 is a circuit schematic for an attenuator. The example is a 'pi' style unbalanced type, such as a 50 \(\Omega\) coaxial device but it could be any type of correctly matched attenuator, balanced or unbalanced.

Figure 4.6-1 A schematic for an pi attenuator with a linear loss \(L\). An input AWGN power of \(N_i \, watts \) is applied at the input. The power absorbed by the attenuator (red arrows) is equivalent to the excess noise power.

The attenuator must be well matched to adjacent stages and in thermal equilibrium at the standard noise temperature \(T_0 = 290 \,\, K\). Before proceeding, we must be familiar with the definitions used for linear and logarithmic gains and losses. These are given in Section 2 under 'Linear and Logarithmic Units'.

The equations will continue to use symbols expressing linear values. An exception is the common attenuator description found in datasheets, or logarithmic loss such as 3 \(dB\). The linear (upper case 'l', unitless) and logarithmic (lower case 'l', dB) values for loss are related by: \(L=10^{l/10}\). For example, a 3 \(dB\) attenuator has a linear loss of 2.0.

Referring to Figure 4.6-1, a noise power of \(N_i \, W\) applied to the input port will be attenuated by linear value \(L\) and then exit the output port. As there is no power supply, by the conservation of energy, the power that was removed to 'create' the attenuation is equivalent to the excess noise power \(N_x\). As we have specified thermally stabilised conditions, we are actually using conservation of power (4.6-1).

\[ N_x = N_i - N_o = N_i - \frac{N_i}{L} = N_i \left(1 - \frac{1}{L} \right) \]
(4.6-1) The excess noise \(N_x\) of an attenuator is the power dissipated to create the attenuation.

The next step is to express the attenuator noise factor \(F\) in terms of excess noise and its loss, very similar to the amplifier case (4.4-3) but here substituting \(G=1/L\) .

\[ F = \frac{N_o}{G N_i} = \frac{G N_i + N_x}{G N_i} = 1 + \frac{N_x}{G N_i} = 1 + \frac{L N_x}{ N_i} \]
(4.6-2) Noise factor of an attenuator with loss \(L\).

Then substituting for \(N_x\) from (4.6-1) into (4.6-2) gives the result for \(F\) (4.6-3).

\[ F = 1 + \frac{L}{N_i} \left[N_i \left( 1 - \frac{1}{L} \right) \right] = L \]
(4.6-3) A matched attenuator of linear loss \(L\) has a noise factor \(F=L\).

Converting to logarithmic units which may be more convenient, the loss of an attenuator in \(dB\) is equivalent to its noise figure in \(dB\).

AWGN in Cascaded Systems

Cascaded 2-port stages are used extensively in communications and radar systems. Their use in receiving systems is of particular interest relating to AWGN because of the desire for low noise. Some reasons for the cascade architecture are listed below.

A very useful approach for cascaded AWGN problems is to start with the AWGN schematic as shown in Figure 4.1-2, then convert this to an equivalent which comprises a hypothetical noise-free DUT with a separate AWGN source at the input represented by \(T_e\). The gain of the DUT remains unchanged. This procedure is shown in Figure 5-1 with the help of (4.4-3) and (4.4-5).

Figure 5-1 A real, noisy DUT with noise factor \(F_A\) (a) may be represented by a noise-free DUT (b) (\(F_B=1\)) with the equivalent input noise temperature \(T_e\).

Using the signal and noise equations that were described in Section 4, Figure 5-1(a) shows how a DUT with a noise factor of \(F_A\) and linear gain \(G\) may be represented. Here, the AWGN source connected to the input represents that used for the noise factor measurement using the \(T_0\) = 290 \(K\) standard. This does not contribute in any way to the internal noise of the DUT, other than to how it is defined.

Figure 5-1(b) is equivalent to (a) but with the internal DUT noise expressed by \(T_e\) referred to the input, obtained from (4.4-5). This schematic shows how this noisy device may be included into a cascade. Moving \(T_e\) values to various input and output positions of the DUTs within a cascade simplifies \(SNR\) related calculations.

To investigate how the \(SNR\) is affected using the same DUT used in Figure 5-1 but with sources of different AWGN properties, refer to Figure 5-2.

Figure 5-2 Input noisy devices of 290 \(K\) (c) and 150 \(K\) (d) may be added to the equivalent schematic in Figure 5-1 (b) to calculate the resulting input and output \(SNRs\).

One advantage of representing the DUT as shown in Figure 5-1 (b) is that its gain is unaffected by any of the noise manipulations. Therefore, the position of \(T_e\) from any device within the cascade may be modified by the appropriate gain factor(s) to other position(s) within a cascade. The derivation of the Friis cascaded stages formula in Section 5.1 demonstrates this principle [3] [13] [19].

Friis's Formula for Cascaded Stages

Now we will derive Friis's cascaded noise factor equation which is very useful for understanding the small signal noise performance of cascaded stages. Again, the assumptions are that each device is operating linearly and well matched to the next and all devices are in thermal equilibrium at the same temperature as the source, which we will take as being the standard noise temperature of 290 \(K\). This approach applies to an input AWGN also at 290 \(K\), such as a nearby signal generator. If the input is a 'colder' source, such as a satellite antenna, we need to take the equivalent input noise temperature approach, which will be described. Refer to Figure 5.1-1.

Figure 5.1-1 The steps used to develop Friis's equation for two cascaded devices under small signal conditions for AWGN.

Figure 5.1-1(a) shows two cascaded stages: 1 and 2 ordered in the direction of the signal-flow. For each, the noise factor and gain symbols are \(F_n\) and \(G_n\) respectively where \(n\) is the stage number. The schematic symbols for amplifiers are used simply for convenience, the same theory applies equally to any other linear 2-port matched devices: attenuators, filters, couplers etc..

Figure 5.1-1(b) is the result of connecting together stages 1 and 2 with each device represented by an equivalent noise-free (\(F=1\)) DUT with an input noise temperature added of the form \(T_{en}\), calculated from (4.4-5). The gains are unchanged and have identical effects on the signal and noise powers as was described in Section 5 and Figure 5-1. In Figure 5.1-1(c), \(T_{e2}\) has been transferred from the input of stage 2 to the input of stage 1, at the same time correcting the value by the reciprocal of \(G_1\), \(1/G_1\). As the noise temperatures are linear functions of AWGN power, they may simply be added as shown in Figure 5.1-1 from (b) to (d).

Assuming the equivalent device for the two cascaded stages has a total gain and total noise factor of \(G_{2T}\) and \(F_{2T}\) respectively, then \(G_{2T}\) is simply the product of each gain separately (5-1).

\[ G_{2T} = G_1 \, G_2 \]
(5-1) The total gain is simply the product of the individual stage gains.

Then (4.4-5) is applied to the schematic representing the \(T_{e}\) value for the equivalent total equivalent input temperature \(T_{e2T}\) (5-2).

\[ F_{2T} = 1 + \frac{T_{e2T}}{T_0} \]
(5-2) (4.4-5) applied to the equivalent (total) single stage to represent 2 cascaded stages.

From Figure 5.1-1(d) the expression for moving the individual stage \(T_e\) values to the input is (5-3)

\[ T_{e2T} = T_{e1} + \frac{T_{e2}}{G_1} \]
(5-3) Provided they are referred to the same position in the cascade, the equivalent noise temperatures may be added at that point.

Then substituting the \(T_{e1}\) and \(T_{e2}\) values from Figure 5.1-1(b), (5-4) and (5-5), into (5-3) gives (5-6).

\[ T_{e1} = T_0(F_1-1) \]
(5-4) From (4.4-5), the equivalent noise temperature for stage 1.
\[ T_{e2} = T_0(F_2-1) \]
(5-5) From (4.4-5), the equivalent noise temperature for stage 2.
\[ F_{2T} = F_1+ \frac{F_2-1}{G_1} \]
(5-6) The resulting expression for the noise factor of 2 cascaded stages in terms of their individual noise factors and the gain of the first stage.

It will be clear from the procedure described for 2 cascaded stages: \(G_{2T}\) (5-1), \(F_{2T}\) (5-2) and \(T_{e2T}\) (5-3), that these may be simply extended to as many stages as required by a linear progression. For example, for 3 stages using the same symbol conventions, \(F_{3T}\), \(T_{e3T}\) and \(F_{3T}\) are given by (5-7), (5-8) and (5-9).

\[ F_{3T} = 1+ \frac{T_{e3T}}{T_0} \, + \, ... \]
(5-7) The total noise factor for 3 stages as a function of the total equivalent input noise temperature and the standard noise temperature.
\[ T_{e3T} = T_{e1} + \frac{T_{e2}}{G_1} \, + \frac{T_{e3}}{G_1 G_2} \, + \, ... \]
(5-8) The cascade noise temperature relationship for 3 stages.
\[ F_{3T} = F_1 + \frac{F_2-1}{G_1} \, + \frac{F_3-1}{G_1 G_2} \, + \, ... \]
(5-9) The cascade noise factor relationship for 3 stages.

From (5-6) we can see that \(F_{2T}\), the overall noise factor of the cascade, in this case only 2 stages, is a significant function of the properties of stage 1 (\(F_1\) and \(G_1\)). In order to minimise the degradation of \(SNR\) caused by the cascade, \(F_1\) must be minimized whilst \(G_1\) must be maximized. Most well designed low noise receive systems therefore include a front-end (stage 1) device with a small noise factor and appreciable gain across the operating frequency band.

Taking the cascade a stage further, it is well known and proved in Section 4.6 that the noise figure of a matched attenuator is equivalent to its logarithmic loss. An interesting exercise using (5-9) is to add such an attenuator to the front end of the cascade and confirm that its effect is to increase the total noise figure of the whole cascade by the same amount as the attenuator loss. Also, the total gain of the cascade is reduced by the same amount, therefore causing two contributions to the degradation of \(SNR\). Therefore the temptation of using a long, relatively high loss, feeder cable from a distant receiving antenna to a receiver may seriously degrade performance.

Most receive systems will comprise many stages and groups of contiguous stages may be analyzed successively. This is usually performed starting at stage 1 because (5-3) and (5-6) show that this is the most critical for AWGN performance. Stage 2 and subsequent stages become progressively less critical.

Noise Measure

Further to Section 5.1 a common question is: "How do I trade off the gains and noise factors of my proposed front end devices for the best noise performance?" One way is to consider 2 front end candidate stages, such as those shown in Figure 5.1-1(a) with the noise factors (\(F_1\) and \(F_2\)) and gains (\(G_1\) and \(G_2\)) as shown. Now consider that the symbol subscripts refer, not to the stage positions, but to the physical devices: numbers, 1 and 2. Suppose, for the order shown in Figure 5.1-1(a), the overall noise factor for the 2 stages is \(F_{TA}\) and for the reverse order (stage 2 followed by stage 1) it is \(F_{TB}\). Then, by applying (5-6) in both cases results in (5.2-1) and (5.2-2) [13].

\[ F_{TA} = F_1+ \frac{F_2-1}{G_1} \]
(5.2-1) The total noise factor for 2 stages in the order: device 1, device 2.
\[ F_{TB} = F_2+ \frac{F_1-1}{G_2} \]
(5.2-2) The total noise factor for 2 stages in the order: device 2, device 1.

If we had assumed \(F_{TA} \lt F_{TB} \), then, using (5.2-1) and (5.2-2) creates the following inequality (5.2-3).

\[ \begin{align} F_{TA} &\lt F_{TB} \\ F_1 + \frac{F_2-1}{G_1} &\lt F_2 + \frac{F_1-1}{G_2} \\ (F_1 - 1) + \frac{F_2-1}{G_1} &\lt (F_2 - 1) + \frac{F_1-1}{G_2} \\ \frac{F_1-1}{1 - \dfrac{1}{G_1}} &\lt \frac{F_2-1}{1 - \dfrac{1}{G_2}} \\ \end{align} \]
(5.2-3) The total noise factors resulting from the alternative orders of device 1 and device 2 used to derive an expression for the noise measure.

The noise measure (\(M\)) of a 2-port device is defined as [13]:

\[ M = \frac{F-1}{1 - \dfrac{1}{G}} \]
(5.2-3) The noise measure of a device with noise factor \(F\) and gain \(G\).

Comparing the noise measure of each of two candidate 2-port devices may therefore be used to determine which device should be first in the signal path to provide the best overall noise performance.

Antenna Noise Temperature

Antenna (equivalent) noise temperature is less a function of the antenna electrical design and more a function of 'where it is pointing', the 'sharpness' of the main lobe and the degradation caused by unwanted side lobes and back lobes. To examine the causes we need to make some assumptions which are listed below.

Measuring Noise Factor Using the Y-Factor Method

The Y-factor method is commonly used by NFMs and NFAs to provide a real-time readout of noise factor (and noise gain) after a suitable calibration procedure. For a DUT without frequency conversion operating within the frequency ranges of the noise source and the NFM, the calibration and measurement steps are shown schematically in Figure 5.4-1 [1] [17] [19].

Figure 5.4-1 Schematics for the calibration and measurement of a DUT without frequency conversion.

The noise source is actually two AWGN sources, either of which may be switched to the output of the device, say at equivalent 'hot' and 'cold' noise temperatures of \(T_H\) and \(T_C\). Often, \(T_C\) is designed to be close to the local ambient temperature. To simplify calculations, it may be convenient to stabilise the ambient temperature to \(T_C = T_0 =290 \, K\) as this is frequently used in datasheets.

'Hot' and 'cold' refer to the greater and smaller equivalent noise temperatures of the noise source state respectively, selected by the 28 \(V\) on-off waveform which originates from the NFM. With a stable ambient temperature of 290 \(K\), then \(T_C=T_0 = 290 \,K\). This is a common test configuration which simplifies the ENR equation and will be used here. Applying this (3-7) for AWGN spectral noise power density (NPD), with \(T = T_0 =290\,K\), and converting to logarithmic units gives us -174 \(dBm/Hz\). To determine \(T_H\) for the same source, (5.4-1) is used with \(T_C=T_0 = 290 \,K\) and obtaining the value for \(ENR_{dB}\) from the noise source datasheet. For example, a common nominal \(ENR_{dB}\) is 14 \(dB\). Using this value and some manipulation of (5.4-1) gives us \(T_H = 7574 \, K\). Applying (3-7) again, but for this temperature, provides the NPD in this case of -160 \(dBm/Hz\), as expected, differing by \(ENR_{dB}\) from the \(T_0\) NPD.

\[ ENR_{dB} = 10 \, log_{10} \left( \dfrac{T_H-T_C}{T_0} \right)\,\,dB \]
(5.4-1) The logarithmic excess noise ratio \(ENR_{dB}\) of a noise source.

The HP 346B noise source has quite a wide operating frequency range of 10 \(MHz\) to 18 \(GHz\). The precise ENR will typically vary by a few fractions of a dB across this range so an ENR against frequency calibration table is included with the device. Data from this may be programmed into the NFM before the measurement calibration to make the necessary minor adjustments. With the exception of the ENR table, all of the remaining discussion on Y-factor measurements will use linear units.

Y-factor is defined as the unit-less ratio of the power measured by the NFM in the \(T_H\) state (\(N_H\,\,watts\)) to the power it measures in the \(T_C\) state (\(N_C\,\, watts\)) (5.4-2). As the NFM itself controls the switching of the sources within the NS, it also stores each value separately in memory for later processing.

\[ Y = \frac{N_H}{N_C} \]
(5.4-2) Y-factor is unit-less and is defined as the ratio of \(N_H\) to \(N_C\).

We know from the beginning of this section that we can replace a noisy device with a calculated equivalent input noise temperature and use a noise-free device with identical gain. This is the principle adopted for the Y-factor analysis. Therefore, with the (switched) noise source connected to the input of the DUT, the measured Y-factor of the DUT is given by (5.4-3) [1] [17] [19]:

\[ Y = \frac{N_H}{N_C} = \frac{G k (T_H +T_e)B_N}{G k (T_C +T_e)B_N} =\frac{T_H +T_e}{T_C +T_e} \]
(5.4-3) Y-factor expressed as the ratio of the 'hot' and 'cold' noise powers measured at the output of a DUT.

Re-arranging (5.4-3) in terms of \(T_e\) gives (5.4-4).

\[ T_e = \frac{T_H-YT_C}{Y-1} \]
(5.4-4) The equivalent input noise temperature in terms of the hot and cold noise temperatures and the Y-factor.

Calibration

Refer to the calibration schematic in Figure 5.4-1. For this example, the (Y-factor) noise factor measurement of a 2-port DUT without frequency conversion will be described. The ENR calibration data across the operating frequency range for the noise source to be used must be stored for processing in the NFM memory. The DUT has a frequency range of 500 \(MHz\) to 1 \(GHz\). This must be within range of both the noise source and the NFM RF input. The equations used for the equivalent noise temperature and noise factor for cascaded stages are shown in terms of noise temperature and noise factor in (5-8) and (5-9) respectively. Before the calibration, the NFM must be set to suitable start, stop and frequency steps and the other parameters set up according to the operating instructions [16].

Calibration must be performed with all of the necessary cables, connectors and adaptors, that will be required to connect the DUT output to the NFM input for the measurement, but excluding the DUT itself. The shortest possible cables and fewest adaptors connecting the noise source to the DUT after calibration must be allowed for. Ideally, the noise source connector type should be the same as the DUT input connector but of the opposite gender. If extra DUT input cables/connectors are unavoidable, their combined loss can be measured and then accounted for by entering the figures into the NFM under 'input loss compensation'.

After performing the calibration, the NFM will store the necessary data and indicate when calibration is complete. This procedure uses the Y-factor (\(Y_{2}\)) to measure the 'second stage' equivalent input noise temperature (\(T_{e2}\)) using (5.4-3), (5.4-4).

\[ T_{e2} = \frac{T_H-Y_2 T_C}{Y_2-1} \]
(5.4-5) Second stage equivalent input noise temperature in terms of Y-factor \(Y_2\), \(T_H\) and \(T_C\)

Furthermore, the NFM controls the switching of the noise source so it 'knows' when it is measuring \(N_H\) or \(N_C\) and can save them as required across frequency for later processing.

Measurement

After calibration, whilst still stabilised, the DUT is connected as shown under 'measurement' in Figure 5.4-1. The same cables and adaptor arrangement that was used for the calibration must now be used for the DUT measurement.

The NFM will now measure the 'hot' and 'cold' noise powers for the cascaded stages 1 and 2 combined: the DUT (stage 1) and the remaining stages including cables, connectors and the devices within the NFM itself (stage 2). We will assume these powers are \(N_{C12}\) and \(N_{H12}\) respectively, then the gain of the DUT, the first stage only, \(G_1\) is given by (5.4-6):

\[ G_1 = \frac{N_{H12}-N_{C12}}{N_{H2}-N_{C2}} \]
(5.4-6) The DUT gain (\(G_1\)) using hot and cold noise powers measured during the Y-factor measurement.

Values for \(N_{C12}\) and \(N_{H12}\) also allow the instrument to calculate the Y-factor for the cascaded stages 1 and 2, \(Y_{12}\), similarly to (5.4-5), (5.4-6)

\[ T_{e12} = \frac{T_H-Y_{12} T_C}{Y_{12}-1} \]
(5.4-6) Substitution for \(Y_{12}\) in (5.4-5).

The final step is then to re-arrange the equation for the noise temperature of two cascaded stages (5-3) in terms of the noise temperature of stage 1, the DUT (5.4-7)

\[ T_{e1} = T_{e12} - \frac{T_{e2}}{G_{1}} \]
(5.4-7) The final result for the equivalent input noise temperature of the DUT (\(T_{e1}\)) using the Y-factor method.

Thus the NFM is then able to calculate the equivalent input noise temperature of the DUT \(T_{e1}\) using the result for the gain of the DUT (\(G_1\)) from (5.4-6). Since we have used the standard noise temperature as a reference, the NFM may readily convert this to noise factor using (4.4-5).

Frequency Conversion

Modern, high specification and very expensive, noise figure analyzers based on real-time digital signal processing (DSP), can perform measurements over many frequency ranges, bandwidths and modes of operation to high accuracy. However, our discussion is based on the Hewlett Packard (HP) 346B Noise Source and the HP 8970B Noise Figure Meter (NFM) [16]. This equipment was designed using the earlier superheterodyne architecture. The HP 8970B processing and download speeds are slower than those available with the DSP equipment but both generations still use the Y-Factor method that is described in Section 5.4.

DUT Without Internal Frequency Conversion

The AWGN-based noise measurements to be described use the Y-factor method which is discussed in Section 5.4. Further to the discussion in Section 2 under 'Linear and Logarithmic Units', the parameters noise figure, noise factor and input noise temperature are synonymous. It is straightforward to convert between them. We will start with a simple noise figure measurement for a DUT without internal frequency conversion which is shown in Figure 5.5-1. The noise measurements must be performed under thermally stabilized conditions at an ambient temperature of 290 \(K\). The frequency range of the DUT, in this example 500 \(MHz\) to 1 \(GHz\), must be within range of both the noise source and the NFM RF input.

The NFM is designed to a low noise specification for measuring AWGN related parameters (noise figure, noise factor, equivalent input noise temperature and noise gain). It is possible for the HP 8970B to measure its own inherent noise figure using a suitable noise source, such as the HP 346B. I have measured this to be within the range 4 dB to 6 dB across 100 \(MHz\) to 1 \(GHz\). The instrument architecture is based upon a superhetrodyne design similar to a traditional (non-DSP) swept frequency spectrum analyzer. However the noise figures of spectrum analyzers are usually much greater than those of dedicated NFMs.

Figure 5.5-1 The calibrate and measure configurations for a DUT without internal (or external) frequency conversion.

If the DUT operating frequency range extends beyond that of the NFM input, but still within the range of the noise source (for the HP346B this is 10 \(MHz\) to 18 \(GHz\)), then frequency conversion is required within the 'second stage' to bring it back to that range. This is frequency conversion, but external to the DUT. One example of the calibrate and measure configurations for a DUT frequency range of 4 \(GHz\) to 5 \(GHz\) with external frequency conversion is shown in Figure 5.5-2.

Figure 5.5-2 Possible calibrate and measure configurations for a DUT using external frequency conversion and a fixed frequency local oscillator (LO).

The mixer and fixed frequency local oscillator (LO) at 3.9 \(GHz\) will down-convert the DUT frequency band to the range 100 \(MHz\) to 1100 \(MHz\), suitable to connect to the NFM RF input. The mixer will generate another product of the noise source in the range 7.9 \(GHz\) to 8.9 \(GHz\) which is rejected from entering the NFM by the low pass filter (LPF) with a cutoff frequency of about 1700 \(MHz\). This may not be essential due to the input filtering of the NFM but is considered good practice. The 'second stage' now includes extra lossy parts (the LPF and mixer) before the NFM which will increase its noise figure compared to the non-internal frequency conversion case. There are some options for mitigating this which are described in the HP8970B user manual [16]. The NFM may be set up to include and compensate for suitable frequency offsets generated by the mixing process. In this case the LO is at a fixed frequency so the NFM frequency stepping capability is used to measure at frequencies across the DUT operating band.

An alternative to the fixed LO configuration shown in Figure 5.5-2 is the variable LO configuration which is shown in Figure 5.5-3

Figure 5.5-3 Possible calibrate and measure configurations for a DUT using external frequency conversion and a variable frequency local oscillator (LO).

As an example, the LO may be tuned from 2.9 \(GHz\) to 3.9 \(GHz\) to give an IF output into the NFM at 1100 \(MHz\). The other mixing product would range from 6.9 \(GHz\) to 8.9 \(GHz\) so a similar LPF to that used for the fixed LO would be advisable. The LO frequency may be controlled either with an automatic test equipment (ATE) configuration or from the NFM itself if the facility is available. This method does have the advantage of allowing the NFM to be tuned to a lower IF such as 100 \(MHz\) where its noise performance may be better than closer to its upper frequency, 1600 \(MHz\).

DUT With Internal Frequency Conversion

Examples of DUTs with internal frequency conversion include mixers, receivers, upconverters and downconverters, in fact any 2-port device in which a swept input frequency does not produce an identical swept output frequency.

There are various options for measuring the noise figures of frequency converting DUTs. Knowledge of the DUT's intended operation, architecture and frequency planning is required to design an optimum (useful and reliable) test method. The examples shown are again featuring the HP 346B noise source and the HP 8970B NFM using the Y-factor method. For both instruments, extensive information is available online including theory, operating instructions, remote operation and examples [16].

Examples of calibration and some example measurement configurations for DUTs with internal frequency conversion are shown Figure 5.5-4 (1) to (4).

Figure 5.5-4 Example configurations for calibration and measurement of a DUT with internal frequency conversion.

Low Noise Antenna Receiving System

An example of a low noise cascaded receiving system devised by Pozar is shown schematically in Figure 6-1 [1] [2].

Figure 6-1 An example low noise front end of a low noise receiver by Pozar [2].

The local oscillator feeding the mixer is not shown. The first 3 stages of the receiver comprise a low noise amplifier (LNA), a bandpass filter and a mixer, each stage with the logarithmic gain (\(g\)), loss (\(l\)) and noise figure (\(f\)) shown in \(dB\). Each symbol has an appropriate subscript. An antenna with an input noise temperature of \(T_A = 150 \,K\), feeds the LNA input. The linear input signal and noise powers are \(S_i\) and \(N_i\) in watts. The linear output signal and noise powers are \(S_o\) and \(N_o\) in watts. The noise bandwidth of the cascade is determined by the passband response of the bandpass filter at 10 \(MHz\). All devices are well matched to the characteristic impedance which is 50 \(\Omega\) and are in thermal equilibrium at \(T_0 = 290 \, K\). The device noise figures were measured with an input temperature also at \(T_0\). The minimum SNR required at the output for the required quality of service is 20 \(dB\).

The second stage, bandpass filter, in Figure 6-1, has an in-band loss of 1 \(d B\). That is equivalent to a logarithmic in-band gain (say \(g_{F}\)) of -1 \(d B\). As the devices are in thermal equilibrium and well matched to the nominal system impedance of 50 \(\Omega\), the noise figure in \(d B\) will be the same as the in-band loss, just like an attenuator described in Section 4.6, so \(f_F\) = 1 \(dB\). These values will be converted to their linear equivalents using upper case symbols and the same subscripts for use in the equations. In summary, using Figure 6-1 and the equations (2-1), (2-2) and (2-3), we have the following linear values:

These are shown in Table 6-1.

Table 6-1 A summary of the logarithmic and linear gains, noise figures and noise factors for the cascade shown in Figure 6-1.
LNA (Stage 1) BPF (Stage 2) Mixer (Stage 3)
\(g_A \, \, (d B) \) \(G_A\) \(f_A \, \, (d B) \) \(F_A\) \(g_F \, \, (d B) \) \(G_F\) \(f_F \, \, (d B) \) \(F_F\) \(g_M \, \, (d B) \) \(G_M\) \(f_M \, \, (d B) \) \(F_M\)
10 10 2 1.585 -1 0.794 1 1.259 -3 0.501 4 2.512

Under Construction

References

  1. Pozar, David M.; Microwave Engineering - Third Edition; John Wiley & Sons Inc.; pp. 487-500 (noise types, black body radiation (Planck), Rayleigh-Jeans approximation, ultraviolet catastrophe, Y-factor, NF of attenuator, Friis, thermal noise, noise temperature). ISBN  0-471-44878-8.
  2. Pozar (op. cit.); pp. 496-497; Example 10.2: Noise Analysis of a Wireless Receiver.
  3. Pozar (op. cit.); pp. 495-496; noise figure of a cascaded system.
  4. Pozar (op. cit.); pp. 493-494; noise figure definition, noise figure of an attenuator.
  5. Sklar, Bernard; Digital Communications, Fundamentals and Applications, Second Edition; Prentice Hall PTR, Upper Saddle River, New Jersey 07458; pp. 30-33 (noise types, AWGN, thermal noise, Gaussian distribution, central limit theorem, white noise, auto-correlation function). ISBN 0130847887.
  6. Sklar, Bernard (op. cit.);
  7. Sklar (op. cit.) p258 thermal noise to 1012 Hz; p282 (sky noise).
  8. Sklar (op. cit.) pp. 116-119; (AWGN, white noise, variance, \(S N R \) related to \(E_b/N_0\)).
  9. Carlson, A. Bruce, Crilly, Paul B., Rutledge, Janet C.; Communication Systems, Fourth Edition; An Introduction to Signals and Noise in Electrical Communications; McGraw-Hill Higher Education, International Edition; pp. 372-381 (Johnson, Nyquist, thermal noise, central limit theorem, Gaussian, white noise, noise power density, standard, temperature, autocorrelation function, many independent charges, noise bandwidth); ISBN 0-07-112175-7.
  10. You Tube® Video, Central Limit Theorem, independent charges, Gaussian noise.. Professor Iain Collings.
  11. You Tube® Video, What is White Gaussian Noise?, auto-corrrelation function Professor Iain Collings.
  12. Carlson et. al. (op. cit.); p398 bandpass noise, AWGN, additive noise, double-sided spectrum.
  13. Gonzalez, Guillermo; Microwave Transistor Amplifiers, Analysis and Design, Second Edition; Prentice Hall, Upper Saddle River, NJ 07458; pp. 295-298 (thermal noise importance, noise figure definition, cascaded noise figure, noise measure); p480 (thermal/random noise to 1000 GHz). ISBN 0-13-254335-4.
  14. Schwartz, Mischa; Information Transmission, Modulation and Noise, Third Edition; International Student Edition; McGraw-Hill Kogakusha, Ltd.; pp. 611-615 (Gaussian/Normal distribution, PDF); ISBN 0-07-055782-9.
  15. Taub, Schilling; Principles of Communications Systems, International Student Edition; McGraw-Hill International Hill Book Company; pp. 236-240 (Noise sources, white spectrum, noise power spectral density); pp. 246-247 (noise power superposition/additive); pp 253-254 noise bandwidth. ISBN 0-07-085790-3.
  16. HP 8970B Noise Figure Meter Operating Manual; Hewlett-Packard Company Test and Measurement Organization, Santa Clara, CA 95052-8059; P/N 08970-91012; specifications, installation, operation, programming, measurement theory.
  17. Noise Figure Measurement Accuracy: The Y-Factor Method; Keysight Application Note 5952-3706E; Keysight Technologies (2024); p 5 Y-factor theory, noise figure/factor; p 9 antenna noise is 1 port, NF 2-port,
  18. 10 Hints for Making Successful Noise Figure Measurements; Keysight Application Note 5980-0288E (05/07/2024); Keysight Technologies; 1 \(Hz\) bandwidth for DSP, choice of ENR, minimise adapters, cable lengths, mixer slope (NF SSB and DSB), isolator to improve match, avoid non-linearities, radiated susceptibility (screened room).
  19. Fundamentals of RF and Microwave Noise Figure Measurements; p3 how does NF relate to digitally modulated systems, relationships: BER, C/N, S/N, NF; p 4 lower NF or increase Tx power?; Keysight Application Note 5952-8255E, 01/10/2019; reasons for low NF, digital services (BER), shielding, filtering, Friis's work, 290 K (IEEE), Y-factor most common method, small ENR for low noise and gain, NF not a function of modulation, cold source method (high NFs).
  20. NFA X-Series Noise Figure Analyzer, Multi-Touch N8973B, N8974B, N8975B, N8976B; Keysight Usr Manual N8973-90007, Ed. 1, 01/2025; Keysight Technologies; SA mode DSP real time measurement BW to 1 \(Hz\), better: frequency accuracy, processing speed, GPIB, LAN, USB, frequency points.
  21. Stutzman, Warren L., Thiele, Gary A.; Antenna Theory and Design, Third Edition; John Wiley & Sons. Inc.; pp 40-44 near, intermediate and far fields; p 50 (directivity and gain) ISBN 978-0-470-57664-9.
  22. Matlab Antenna Toolbox Users Guide, R2023a; The Mathworks Inc.; pp 1-21 - 1-30 antenna concepts: near/far fields, directivity, gain, beamwidth, E and H planes, polarization.
  23. Stutzman et. al. (op. cit.); pp 103-107Antenna noise temperature related to elevation.
  24. Kraus, J. D. and Carver, K. R.; Electromagnetics, Second Edition; McGraw-Hill, Electrical and Electronic Engineerinf Series; ISBN 0-07-035396-4; p13 inverse square law, spherical spreading.
  25. Stutzman et. al. (op. cit.); p3, p109, p395, p421 inverse square law, spherical spreading
  26. Kraus, J. D. and Carver, K. R. (op. cit.); pp. 606-617; near, far and intermediate field distances.
  27. Amphenol-Andrew Antenna; parabolic reflector antenna part HX6-11W-2WH: d=1.8 m, 10-11.7 GHz, gain = 43 dBi, beamwidth = \(1^\circ \times 1^\circ\), XPD = 33 dB, F/B = 76 dB.