Direct-coupled amplifier
A direct-coupled amplifier, also known as a DC amplifier, is an electronic circuit that provides useful voltage gain at zero frequency by connecting the output of one amplification stage directly to the input of the next without intermediate coupling capacitors or other frequency-selective components.[1] This design enables the amplification of signals with significant direct-current (DC) components or very low frequencies, down to fractions of a hertz, making it essential for applications involving steady-state or slowly varying inputs.[2]
The operating principle relies on active devices like bipolar junction transistors or vacuum tubes biased to handle DC levels, often incorporating differential amplifier configurations to minimize unwanted offsets and ensure stable performance across stages.[1] Unlike AC-coupled amplifiers, which use capacitors to block DC and limit low-frequency response, direct-coupled designs avoid such elements, providing a flat frequency response from DC upward but introducing challenges like thermal drift—unintended output variations due to temperature changes in transistor parameters such as base-emitter voltage (V_BE ≈ 0.6 V) or current gain (β).[2] Advantages include simplicity, low cost, minimal component count, and excellent low-frequency performance without the need for large capacitors or long recovery times after transients, while disadvantages encompass sensitivity to power supply variations, aging effects, and drift that can limit the minimum detectable signal to microvolts or worse in uncompensated circuits.[1][2]
Historically, direct-coupled amplifiers were challenging to implement reliably in the mid-20th century due to drift issues, as noted in early texts advising against their construction without advanced stabilization techniques; however, advancements in integrated circuits and feedback methods have made them foundational to modern operational amplifiers and precision instrumentation.[1] Key applications span biological signal processing (e.g., electrocardiograms), geological measurements, voltage regulators, and high-speed data links, where DC-40 GHz bandwidths have been achieved in monolithic microwave integrated circuits (MMICs) using heterojunction bipolar transistors.[2][3] In operational amplifiers, they form the input stage to handle DC biases while achieving high common-mode rejection ratios (up to 10^6) and low input offsets through matched transistor pairs.[2]
Fundamentals
Definition and Operating Principles
A direct-coupled amplifier consists of multiple amplifier stages interconnected without intermediate coupling components such as capacitors or transformers, relying instead on direct electrical connections via resistive or active elements to transmit both alternating current (AC) and direct current (DC) signals between stages.[4] This configuration, also known as a DC amplifier, enables the amplification of signals across a frequency range extending from zero hertz (DC) to high frequencies, as there is no low-frequency roll-off introduced by coupling capacitors.[4]
As foundational knowledge, amplifier stages in such circuits typically provide voltage or current amplification using active devices like bipolar junction transistors (BJTs) or vacuum tubes, where each stage processes the input signal to produce an amplified output while maintaining signal integrity.[5] In a multi-stage direct-coupled setup, the output of one stage connects directly to the input of the next, forming a chain that preserves DC operating points across stages; a basic block diagram illustrates this as follows:
Input Signal → [Amplifier Stage 1] → Direct Connection → [Amplifier Stage 2] → Direct Connection → ... → [Output Stage] → Output Signal
Input Signal → [Amplifier Stage 1] → Direct Connection → [Amplifier Stage 2] → Direct Connection → ... → [Output Stage] → Output Signal
This direct linkage ensures that DC bias levels propagate through the amplifier, allowing the entire system to respond to static (DC) inputs as well as dynamic (AC) variations.[6]
The operating principle hinges on the interdependent biasing of stages, where the DC output voltage of a preceding stage sets the quiescent point for the subsequent one, facilitating broadband response without frequency-selective filtering from coupling elements.[4] Early implementations appeared in vacuum tube circuits during the 1930s, addressing the need for DC amplification in applications like instrumentation.[7]
For a simple direct-coupled stage using a common-emitter BJT configuration with an unbypassed emitter resistor R_e, the voltage gain A_v can be derived from the transistor's small-signal characteristics as follows. Consider the small-signal model where the BJT acts as a transconductance amplifier with g_m = I_C / V_T (where I_C is the collector quiescent current and V_T is the thermal voltage, approximately 26 mV at room temperature), and the intrinsic emitter resistance r_e = 1 / g_m. With an input voltage v_{in} applied to the base and the emitter connected through R_e to ground, the emitter current i_e is approximately v_{in} / (r_e + R_e), neglecting base current for high current gain \beta. The collector current i_c \approx i_e, and the output voltage v_{out} = -i_c R_c, where R_c is the collector resistor. Thus, A_v = v_{out} / v_{in} = -R_c / (r_e + R_e). When R_e \gg r_e, this approximates to A_v = -R_c / R_e, providing stable gain independent of transistor variations.[5]
Comparison to Other Coupling Methods
Direct-coupled amplifiers differ from traditional coupling methods primarily in their ability to handle DC signals and achieve a broader frequency response, though they introduce challenges in biasing stability. In capacitive coupling, also known as RC coupling, capacitors are employed to block DC components while allowing AC signals to pass between amplifier stages, thereby isolating the biasing of each stage. This method results in a high-pass filter characteristic in the frequency response, with a low-frequency roll-off determined by the capacitor's reactance, typically limiting effective amplification below 50 Hz and preventing the amplification of DC signals altogether.[8][9]
Transformer coupling, on the other hand, utilizes a transformer to interconnect stages, providing excellent impedance matching and efficient power transfer, particularly in output stages. However, it imposes frequency limitations due to the transformer's inductive reactance and core saturation effects, leading to poor low-frequency response and a band-pass-like characteristic that excludes DC amplification. Transformers also contribute to increased size, weight, and cost compared to other methods, making them less suitable for compact, low-frequency applications.[8][9]
In contrast, direct coupling connects stages without intervening reactive components, enabling a flat frequency response extending to 0 Hz and preserving DC accuracy for applications requiring signal fidelity from DC to high frequencies. This approach avoids the phase shifts at low frequencies inherent in capacitive and transformer methods, which arise from their reactive elements, but necessitates precise biasing techniques to mitigate thermal drift and offset accumulation across stages. Qualitatively, the frequency response of a direct-coupled amplifier remains constant across the spectrum, unlike the high-pass roll-off in capacitive coupling or the narrower bandwidth in transformer coupling.[8][9]
| Coupling Method | DC Handling | Frequency Response Characteristics | Key Trade-offs |
|---|
| Capacitive (RC) | Blocks DC | High-pass filter; roll-off below ~50 Hz | Simple and low-cost, but limited low-frequency extension and no DC amplification[8][9] |
| Transformer | Blocks DC | Band-pass; poor at low frequencies due to core effects | Excellent impedance matching, but bulky and costly with narrow bandwidth[8][9] |
| Direct | Passes DC | Flat from 0 Hz to high frequencies | Broadest response and DC preservation, but requires precise biasing to avoid drift[8][9] |
Historical Development
Origins and Early Concepts
The conceptual roots of direct-coupled amplifiers emerged in the 1920s and 1930s during the vacuum tube era, as engineers sought to extend amplification beyond the limitations of AC-coupled designs prevalent in early radio receivers, which blocked DC signals and restricted low-frequency response.[10] This evolution was particularly driven by the demand for DC amplification in applications such as cathode-ray oscilloscopes, where visualizing steady-state voltages alongside transient signals required circuits capable of passing DC without capacitors or transformers.[11]
Early concepts focused on direct-coupled multi-stage vacuum tube configurations to achieve broadband response, but faced significant hurdles from tube characteristics. In 1930, Edward H. Loftin and S. Y. White introduced a pioneering two-stage direct-coupled amplifier using triodes like the 245, where the plate of the driver tube connected directly to the grid of the output tube via a resistor, enabling wide frequency coverage from audio lows to highs without interstage coupling elements.[7] Their design, detailed in Radio News articles, addressed grid current distortion by cascading valves with one tube's grid-cathode path in series with the next, as patented in British Patent No. 362,419.[10] However, practical implementation was limited by vacuum tube variability, including inconsistent emission and filament aging, which caused drift and instability in DC paths.[12]
A key advancement came in 1927 when H.S. Black at Bell Laboratories invented negative feedback, patented in 1937 (US Patent 2,102,670), which provided a method to stabilize amplification and counteract tube variations, enabling more reliable direct-coupled designs.[13] These foundational efforts laid the groundwork for DC-responsive amplification, prioritizing conceptual stability over exhaustive low-frequency metrics.
Key Milestones and Inventors
Following World War II, direct-coupled amplifiers gained prominence in military radar systems and early computing applications due to their ability to maintain DC stability for precise signal processing, as evidenced by advancements at the MIT Radiation Laboratory where vacuum tube-based designs were refined for analog computation in gun directors and radar tracking.[12] This era marked a shift toward practical implementation, building on wartime needs for reliable DC amplification in feedback control systems. The invention of the transistor in December 1947 by John Bardeen, Walter Brattain, and William Shockley at Bell Laboratories revolutionized the field, enabling compact, low-power solid-state designs that reduced size and thermal drift compared to bulky vacuum tubes.[12]
In the 1950s, direct-coupled amplifiers became integral to analog computers, where chopper-stabilized variants provided the necessary DC accuracy for solving differential equations in scientific and engineering simulations, as detailed in contemporary works on electronic analog computing.[12] The transition to silicon transistors in the 1960s further mitigated thermal issues inherent in earlier germanium devices, allowing for more stable direct coupling in operational amplifiers. A pivotal milestone occurred in 1964 when Robert (Bob) Widlar at Fairchild Semiconductor designed the μA702, the first monolithic integrated circuit operational amplifier with direct-coupled inputs, which integrated all components on a single chip and set the standard for modern IC designs.[14] This was followed in 1968 by Fairchild's release of the μA741, the first widely commercialized direct-coupled IC op-amp, featuring internal frequency compensation that enhanced usability and reliability.[15]
During the 1970s, James Solomon advanced direct-coupled amplifier performance through innovations in offset and drift reduction, notably in his 1974 tutorial on monolithic op-amps, which introduced transconductance reduction and layout techniques to minimize input errors in bipolar ICs, influencing designs like National Semiconductor's LM324 quad op-amp released in 1972. The 1980s saw further evolution with the rise of CMOS technology, enabling low-power direct-coupled amplifiers suitable for battery-operated and integrated systems, as CMOS processes allowed for rail-to-rail input/output ranges and reduced static power consumption in op-amps.[12]
Design and Implementation
Basic Circuit Configurations
Direct-coupled amplifiers employ straightforward topologies that eliminate coupling capacitors, allowing DC and AC signals to propagate through the circuit without interruption. The simplest single-stage configuration is the common-emitter amplifier using a bipolar junction transistor (BJT), where the input signal is applied to the base terminal, the emitter is grounded through a resistor to provide negative feedback for thermal stability, the collector serves as the output directly connected to the load resistor, and the collector resistor sets the DC operating point while contributing to the AC gain.[16] In this setup, the emitter resistor stabilizes the bias point against variations in transistor parameters, ensuring reliable operation across temperature changes.[16]
An analogous single-stage design for field-effect transistors (FETs) is the source follower (common-drain) configuration, which acts as a unity-gain buffer. Here, the input is fed to the gate, the source directly drives the load with minimal voltage drop, and the drain connects to the positive supply; a source resistor may be included for biasing, but the direct connection to the load preserves low-frequency response down to DC.[16] This topology offers high input impedance and low output impedance, making it suitable for impedance matching in direct-coupled systems.[17]
For higher gain requirements, multi-stage direct-coupled amplifiers cascade multiple BJT or FET stages, often incorporating level-shifting elements like resistors or diodes to align DC operating points between stages without introducing low-frequency roll-off. A common example is a differential input pair—consisting of two matched transistors with tails connected to a current source—directly coupled to a subsequent common-emitter or common-source gain stage, as seen in basic operational amplifier designs.[2] In such cascades, bias currents are shared across stages to maintain overall DC balance.[18]
Specific configurations enhance performance in these topologies. The voltage-follower arrangement, based on an emitter follower for BJTs or source follower for FETs, provides unity gain (A_v = 1) and is inherently direct-coupled, with the output taken from the emitter or source directly following the input voltage.[17] Current-mirror biasing circuits, using matched transistor pairs to replicate a reference current, establish precise DC paths throughout the amplifier, ensuring consistent operating points in multi-stage designs.[16] Additionally, constant-current sources—implemented via transistors or integrated circuits—serve as active loads to minimize loading on prior stages, preserving signal integrity and maximizing gain by presenting a high incremental resistance.[17]
In multi-stage direct-coupled amplifiers, the overall voltage gain A_{total} is the product of the individual stage gains, expressed as
A_{total} = \prod_{i=1}^{n} A_i,
where n is the number of stages; this multiplicative relationship arises because the direct DC connection allows the full output voltage of each stage to drive the input of the next, without capacitive isolation that would block DC components.[18]
Biasing and Stabilization Techniques
In direct-coupled amplifiers, DC biasing establishes the quiescent operating point for transistors across stages without isolation capacitors, ensuring proper signal amplification while preventing saturation or cutoff. Voltage divider networks, formed by two resistors connected from the collector supply to ground and tapped to the base, provide a stable base voltage that is relatively independent of the transistor's current gain (β), mitigating shifts in collector current due to β variations typically ranging from 50 to 300 in bipolar junction transistors (BJTs). This Thevenin-equivalent approach sets the emitter current through a resistor, achieving bias stability better than 10% over β fluctuations. Diodes, such as forward-biased junctions (around 0.7 V for silicon), can provide a reference voltage to clamp the base potential, though they exhibit temperature dependence (approximately -2 mV/°C). For better stability, bandgap references or compensated Zener diodes at higher voltages (e.g., 5-6 V with low temp coeff) are used in precision designs.[19] However, β variations remain a challenge, as even small changes can alter the quiescent current by up to 20-30% without additional stabilization, necessitating careful resistor selection with values like 10 kΩ to 100 kΩ for typical audio or instrumentation stages.
Stabilization techniques address these biases by maintaining the operating point against temperature, aging, and component mismatches. Overall DC feedback loops, implemented by resistors connecting the output to the input (e.g., a 1 MΩ feedback resistor in a multi-stage configuration), sense deviations in the output DC level and adjust the input bias accordingly, reducing offset errors by factors of 100 or more through negative feedback. This method corrects cumulative stage offsets in direct-coupled chains, where a 1 mV error in one stage can amplify to volts at the output. Chopper stabilization enhances precision by using electronic switches to modulate the input signal to a carrier frequency (typically 1-10 kHz), amplify it, demodulate, and subtract the DC offset via an auxiliary nulling amplifier, effectively nulling offsets to below 1 µV and drifts to 0.1 µV/°C. Modern variants may incorporate auto-zero phases for additional offset correction, but classical chopper operation relies on continuous switching, with low-pass filters to suppress the offset and ripple.[20] In practice, anti-aliasing filters may be used to prevent noise folding in switched systems, making it ideal for low-frequency DC amplification without introducing significant ripple.
Additional techniques include thermal tracking via matched devices and precise current sources. Matched transistors, fabricated on the same chip and placed in thermal proximity, exhibit V_BE variations of less than 1 mV/°C, compensating for temperature-induced bias shifts by ensuring uniform thermal coefficients across stages. The Widlar current source, a modified current mirror with an emitter degeneration resistor (typically 1-10 kΩ) in the output branch, generates stable low currents (e.g., 1-100 µA) much smaller than the reference current, improving output impedance to over 1 MΩ and bias precision in tail currents for differential pairs.[21] This configuration solves the limitation of basic mirrors, where output currents cannot be significantly lower than input without mismatch sensitivity.
Mismatches in differential input pairs, common in direct-coupled front ends, introduce offset voltages primarily from ΔV_BE differences (around 1-5 mV due to process variations). The input-referred offset voltage is given by
V_{os} = \frac{\Delta V_{BE}}{1 + \frac{R_f}{R_g}}
where ΔV_BE is the base-emitter voltage mismatch, R_f is the feedback resistor, and R_g is the gain-setting resistor in a non-inverting configuration; this derives from the differential pair's small-signal model, where the offset at the output (A_v ΔV_BE) is divided by the closed-loop gain (1 + R_f / R_g) to refer it to the input. For unmatched pairs, ΔV_BE arises from geometric or doping variations, amplifying to output errors unless compensated by the above techniques.
Advantages and Bandwidth Benefits
Direct-coupled amplifiers provide a wide bandwidth that extends from DC (0 Hz) to frequencies in the GHz range, without the low-frequency cutoff inherent in capacitively or transformer-coupled designs. This full-spectrum response is particularly advantageous for amplifying pulse and square-wave signals, as it preserves the sharp transitions and steady-state levels without distortion from frequency-selective filtering. For instance, GaAs MESFET-based direct-coupled amplifiers have achieved 10 GHz bandwidth with 20 dB gain, demonstrating their capability for high-speed applications.[2][22][23]
A primary strength lies in the high-fidelity amplification of signals containing DC components, such as those encountered in video processing or sensor interfaces, where maintaining absolute signal levels is critical for accuracy. Unlike AC-coupled alternatives, direct coupling ensures no attenuation or phase shift at low frequencies, enabling precise reproduction of baseline voltages and slow-varying waveforms.[24]
In integrated circuit design, direct coupling offers significant advantages by eliminating the need for large on-chip coupling capacitors, which occupy substantial area and increase fabrication costs. This results in simpler topologies with reduced component counts compared to coupled amplifiers, facilitating compact and efficient monolithic implementations.[25]
The upper 3 dB bandwidth f_{3\mathrm{dB}} is determined by the interaction between effective resistance R and parasitic capacitances C_{\mathrm{parasitic}}, approximated as
f_{3\mathrm{dB}} \approx \frac{1}{2\pi R C_{\mathrm{parasitic}}},
where C_{\mathrm{parasitic}} encompasses device junction capacitances (e.g., C_{be}, C_{bc}) and wiring capacitances. To derive this, consider the high-frequency equivalent circuit: parasitic capacitances form a low-pass filter with the Thevenin resistance at the node, creating a dominant pole at \omega_p = 1/(R C_{\mathrm{parasitic}}); the 3 dB point occurs when the gain drops to $1/\sqrt{2} of midband value, yielding the frequency expression. In direct-coupled designs, the absence of intentional coupling capacitors minimizes additional parasitics, often extending bandwidth beyond that of coupled stages.[26]
Direct-coupled amplifiers enable true DC amplification, which is essential for circuits like logarithmic amplifiers and precision rectifiers that demand accurate handling of steady-state and near-DC signals without baseline wander.[25][27]
Drift Mechanisms and Compensation
In direct-coupled amplifiers, drift mechanisms arise primarily from thermal variations and component instabilities, which can significantly degrade performance by causing unintended shifts in the output voltage over time or temperature changes. A key contributor is the temperature coefficient of the base-emitter voltage (V_BE) in bipolar junction transistors (BJTs), typically around -2 mV/°C, which leads to output voltage shifts as the junction temperature fluctuates. This effect is exacerbated in direct-coupled designs due to the absence of coupling capacitors, allowing DC paths to propagate these variations directly through multiple stages. Additionally, aging in resistors introduces gradual resistance changes, while leakage currents in semiconductors can vary with temperature, further amplifying low-frequency errors.
Drift in these amplifiers manifests in two primary types: thermal drift from long-term heating during operation and transient drift occurring immediately after power-on. Thermal drift results from steady-state temperature gradients across the circuit, often leading to output offsets on the order of several mV/°C in simple uncompensated designs, scaling with gain.[2] In contrast, transient drift is driven by initial thermal transients and charge imbalances, which can cause temporary excursions before stabilization. These mechanisms are particularly pronounced in multi-stage direct-coupled amplifiers, where cumulative errors from cascaded devices compound the issue.
To mitigate these drifts, various compensation techniques are employed, focusing on stabilizing the offset and gain against temperature and time. Auto-zero circuits periodically sample and subtract the input offset voltage, effectively nulling thermal-induced errors in real-time. Correlated double sampling (CDS) is another method, where the signal is sampled twice—once with noise and once without—to cancel out low-frequency drifts like 1/f noise and thermal offsets. Modern chopper-stabilized and auto-zero operational amplifiers achieve input-referred thermal drifts below 1 µV/°C.[28] The thermal drift rate can be quantitatively expressed as:
\frac{\Delta V_{out}}{\Delta T} = \left( \frac{\partial A_v}{\partial T} \right) V_{in} + A_v \left( \frac{\partial V_{os}}{\partial T} \right)
where A_v is the amplifier gain, V_{in} is the input voltage, V_{os} is the input offset voltage, and the partial derivatives account for temperature dependencies of gain and offset. This equation highlights how compensation must address both gain and offset variations to minimize output drift.
Applications and Modern Uses
Traditional Audio and Signal Processing
In traditional audio applications, direct-coupled amplifiers have been employed in power stages to enhance subwoofer control and transient response by providing a direct DC-to-audio frequency path without the low-frequency roll-off inherent in capacitor-coupled designs. This configuration is particularly beneficial for driving low-frequency drivers, as it maintains precise cone positioning and rapid response to signal changes, avoiding the phase shifts and compression that can occur with coupling capacitors under high signal excursions. For instance, electrolytic capacitors in the signal path can introduce harmonic distortion when subjected to large AC voltages, as their dielectric properties lead to non-linear behavior, whereas direct coupling eliminates this issue entirely.[29]
A key advantage in these audio power applications is the elimination of the audible "thump" caused by capacitor charging transients during power-on or switching events in capacitor-coupled designs, though direct-coupled amplifiers require mechanisms like DC servos or output relays to manage offset transients and prevent damaging DC pulses to speakers. Direct-coupled designs, such as those used to drive the Quad ESL-57 electrostatic speakers, have been integral since the late 1950s, with later solid-state implementations like the Quad 405 amplifier (introduced in 1976) exemplifying this approach for improved damping and low-end accuracy.[30] These amplifiers ensure stable DC offset control, preventing woofer excursion problems while preserving the full dynamic range of audio signals.
In signal processing chains, direct-coupled amplifiers facilitate full-spectrum handling in mixers and equalizers, enabling seamless transmission of bass-heavy content without attenuation or added noise from interstage capacitors. For example, professional studio consoles from the 1970s and early 1980s, such as the SSL 9000J series, incorporated DC-coupled circuits to remove capacitors from the signal path, resulting in cleaner low-frequency response and reduced phase distortion across the audio band. This design choice contributed to lower overall noise floors in analog mixing environments.
Direct coupling also plays a role in vinyl phono preamplifiers, where it allows the preservation of infrasonic signals from the cartridge without high-pass filtering imposed by coupling capacitors, thereby maintaining the integrity of the RIAA equalization curve down to very low frequencies.[31] However, this necessitates dedicated rumble filters post-preamplification to attenuate subsonic noise from turntable mechanics, preventing woofer pumping while avoiding interference with audible bass content.
Precision Instrumentation and Integrated Circuits
Direct-coupled amplifiers are critical in precision instrumentation systems, such as digital multimeters and thermocouple interfaces, where accurate measurement of low-level DC voltages is essential. In multimeters, these amplifiers form the input stage to handle DC signals from millivolts to volts without introducing offsets or phase shifts that could compromise accuracy, enabling resolutions down to microvolts. For thermocouples, which produce small DC voltages proportional to temperature differences, direct-coupled designs facilitate ratiometric measurements by maintaining a stable reference and avoiding capacitor-induced low-frequency roll-off, thus ensuring offset-free amplification and high common-mode rejection for reliable temperature sensing.[32][33]
In integrated circuit implementations, direct-coupled operational amplifiers like the OPA227 from Texas Instruments are employed in feedback loops for high-precision applications, offering low offset voltage (typically 10 μV) and minimal drift to preserve signal integrity across DC and low-frequency bands.[34] Similarly, the AD797 IC, introduced by Analog Devices in the 1990s, exemplifies advanced direct-coupled design with an input offset voltage drift of less than 1.0 μV/°C, making it suitable for instrumentation where thermal stability is paramount. These ICs integrate direct coupling to eliminate interstage capacitors, reducing parasitic effects and enabling bandwidths from DC to several MHz while maintaining precision in feedback configurations.[35]
Modern applications extend direct-coupled amplifiers to sensor interfaces and biomedical systems. In MEMS accelerometer interfaces, DC-coupled amplification directly processes static and low-frequency acceleration signals, achieving sub-hertz response without baseline wander, as seen in precision vibration monitoring circuits. For biomedical electrocardiogram (ECG) amplifiers, direct-coupled topologies provide high input impedance and common-mode rejection ratios exceeding 100 dB, essential for capturing millivolt-level cardiac signals amid electrode offsets. In the 2020s, cryogenic direct-coupled amplifiers have emerged for quantum computing readout, operating at temperatures near 4 K to amplify weak superconducting nanowire single-photon detector signals with minimal added noise, supporting high-fidelity qubit state discrimination. Recent advancements as of 2024 include hybrid topologies for low-noise direct-coupled front-ends in sensor applications, enhancing voltage and current noise performance.[36][37][38][39]
In precision IC contexts, the noise figure (NF) of direct-coupled amplifiers quantifies signal degradation due to input voltage noise e_n, particularly when thermal noise is secondary. For low signal levels, it is approximated as
NF = 10 \log_{10} \left(1 + \left( \frac{e_n}{V_{sig}} \right)^2 \right)
where e_n is the RMS input-referred voltage noise density integrated over bandwidth, and V_{sig} is the RMS signal voltage; this metric highlights the amplifier's contribution to overall system noise in DC-precise environments like sensor readouts.[40]