Chapter 5 Designing and Structuring Pipelines

When we start writing a new piece of software, one of our first challenges is to identify its logical components and how they interact with each other. We can then structure our software into a series of modules, be they classes, libraries or completely separate programs, that implement those logical components in such a way as to make reasoning about the software as easy as possible. In other words, we design software to divide and conquer complexity into manageable chunks so that we only need to face a small fraction of it at any given time (Ousterhout 2018). Failure to do so quickly leads to software that is impossible to understand and to work on (Chapter 6), which in turn makes it difficult to deploy (Chapter 7), document (Chapter 8), test or troubleshoot (Chapter 9), and in general to keep running.

In this chapter we discuss the unique challenges that define machine learning software design: the role of data (Section 5.1), the nature of technical debt (Section 5.2) and the anatomy of a machine learning pipeline (Section 5.3).

5.1 Data as Code

The inversion of roles in machine learning software (right) compared to other software (left).

Figure 5.1: The inversion of roles in machine learning software (right) compared to other software (left).

Machine learning software is fundamentally different from most other software in one important respect: it is tightly linked with data (Arpteg et al. 2018). The structure and the behaviour of a piece of traditional software11 arise from some combination of processes gleaned from experts in the field, a specification of the desired output, and the set of technologies we can use to support its operations (Figure 5.1, left). We are in charge of designing a software architecture that produces the desired behaviour. For instance, we structure web services to direct user navigation patterns through established procedures for different tasks, taking information retrieved from some database or from various vendor APIs and producing outputs to be consumed through some dashboard (by humans) or API (from other computer systems). Desktop applications do the same through windows and dialogs. Obviously, our freedom in designing software architectures is limited for good reasons (performance requirements, good practices and maintainability among them) as well as bad reasons (like less-than-ideal requirements, limitations in the chosen technological stack and unclear requirements) but this still leaves us a substantial amount of control.

On the other hand, the behaviour of machine learning software is dictated as much by the data we train our models on as it is by our design choices. We may decide how to measure model performance but the best performer will then be determined by the data: the distribution of the variables in the data and their probabilistic structure will be better captured by some models than others. So we may choose to try, say, random forests, deep neural networks and some hierarchical Bayesian model but, in the end, we will end up using the model that the data say is best regardless of our personal preferences. The information in the data is compiled into the software through the models, which program the software automatically: developers do not completely encode its behaviour in the code (Figure 5.1, right).

This realisation leads to a paradigm shift: we should treat data as code because data functionally replaces parts of our source code and because changes in the data may change the behaviour of the software. Hence we should test the data to ensure that their characteristics do not change over time (Section 5.2.1). After all, if the data change, our models may no longer be fit for purpose and we may have to retrain them to retain suitable levels of performance. In the case of offline data, this means that data should be versioned along with the code and that changes in either of them should trigger testing by continuous integration tools. In the case of online data, we should also implement a real-time monitoring and logging of the characteristics of new data and of the performance of the deployed models (Section 5.3.6). Once we are confident that the data are as we expect them to be, we can use them to test that our software (the implementation) is behaving correctly and ensure that the models themselves are correctly specified (in their mathematical and probabilistic formulation). We will discuss the troubleshooting and testing of both data and machine learning models in more detail in the next section and in Chapter 9.

Ideally, we should have a configuration management platform (often called an “experiment tracking” or “experiment management” platform in this context) using version control (Section 6.5) to track the hardware, the source code, the environment configurations, the parameters, the hyperparameters, the model characteristics, the input data and the outputs of all instances of model training and inference. (Including those we use to explore the data.) We can then tag the exact version of all the components used in each development and production environment, as we would do in a traditional software engineering setting. In turn, this means that we can (re)create any of those environments as needed, which makes automated deployments possible (Chapter 7) and greatly facilitates troubleshooting. Given the limited interpretability and explainability of most machine learning models, which are essentially black boxes, only a solution approaching a reproducible build setup (Humble and Farley 2011) can hope to make in-depth debugging and root cause analyses possible.

5.2 Technical Debt

Treating data as code means we should consider data a potential source of technical debt. Models can also be sources of technical debt because of their dependence on data and in their own right. In practice, the data and the models are dependencies of our machine learning code: like all dependencies, they are a potential liability and should be handled as such.

The term “technical debt” has commonly had a negative connotation since it was first introduced (Cunningham 1992, 2011): it highlights how hasty design choices can lead to unexpected costs, not only in purely economic terms, but by introducing latent complexity that makes the software more difficult to evolve over time. Technical debt allows us to produce results faster by trading quality for speed but, as with borrowed money, we must eventually pay it off with (compound) interest. It is unavoidable when tight deadlines reduce the time spent on analysis and design (Evans 2003), leading to solutions that are suboptimal in terms of functionality, code quality or technical implementation. Establishing and following the practices we advocate in Part 2 of this book is a good way of keeping it in check and of paying it off quickly enough to reduce it over time.

Machine learning models and the underlying training, testing and serving software infrastructure, which we will introduce in Section 5.3 as a machine learning pipeline, combine all the complexities of traditional software development with the issues arising from the experimental nature of data analysis. (More about this in Chapter 6.) Therefore, we find it useful to rethink the nature of technical debt in machine learning software in a unified, comprehensive way. We classify it into four broad areas: data, model, architecture (design) and code debt. These areas span issues both in various parts of the machine learning practice, such as data collection, data validation, feature extraction, data visualisation and observability; and in the software that we use to interact with machine learning models, such as monitoring, configurations, training and serving infrastructure. The libraries that power the models themselves, like PyTorch (Paszke et al. 2019) or Scikit-learn (Scikit-learn Developers 2022), are typically very stable and we rarely find them to be a source of technical debt.

5.2.1 At the Data Level

Section 5.1 suggests that data can be a liability for three reasons. Firstly, they may originate from untrusted sources, either from in-house or from third-party systems. Data sources that are outside of our control or that do not have strict quality standards should be treated as an unknown quantity: data may unexpectedly change over time in shape (raw data structure or type change), in general quality (data duplication, missing data, null data or incorrectly normalised data) or in relevance and statistical properties (data or concept drift). This is particularly the case for online data that come in the form of event streams or that are generated by aggregating data from multiple sources. (More on that in Sections 9.1 and 9.4.3.) In order to prevent such anomalies from affecting both the training of models and their subsequent use, we should only allow data that have been versioned and validated by our suite of software tests (Section 9.4) to enter the machine learning pipeline. Systematic testing acts as a quality gate that the data must pass before entering later processing stages. Data drift will make models become stale: their accuracy will decrease as the data they will perform inference on become increasingly different from those that were used to train them (Gama et al. 2014 is an extensive review of this topic). The same may happen if the general quality of the data degrades over time. Unless such changes are sudden enough and sharp enough, their effects will be difficult to detect without a test suite. This is what appears to have happened to Zillow (Sherman 2022), the online real-estate company: the machine learning model they used to price properties to buy was trained on self-reported data, which were untrusted and difficult to validate, and it was left to overestimate prices for too long as the market cooled down. By the time the model was retired in 2021, Zillow had to sell between 60% and 85% of the properties it bought at a loss and fire 25% of its staff just to remain afloat.

Secondly, data may originate from untracked sources: we should always take into account that third-party sources can be volatile and can also suddenly become unavailable. If that happens to a data source we are not aware we depend on, troubleshooting the resulting issues may be challenging. Furthermore, untracked sources are often untrusted as well, but unlike tracked sources they are not systematically versioned and validated: any issue they may have can potentially go unnoticed for long periods of time. In this context, where a piece of data comes from and how it was produced is called data provenance or data lineage (Cheney, Chiticariu, and Tan 209AD).

Finally, we may introduce in the data when we prepare them for use in the pipeline. In many applications, we can only collect unlabelled data that we have to annotate manually: this is an expensive, time-consuming and error-prone process that requires a team of domain experts. Automated labelling using machine learning models is a poor substitute as it is known to have 0.15–0.20 lower accuracy for both natural language processing and computer vision tasks (Wu et al. 2022). The lack of ground truth labels makes it very difficult to spot these errors, which in turn impacts both other data quality controls and model training. Furthermore, manual labelling is too slow to allow us to monitor the outputs of the pipeline in real time, limiting our ability to detect data drift and model staleness. Hence this issue can produce technical debt at different levels in ways that are difficult to detect.

5.2.2 At the Model Level

Issues with model performance, caused by data or otherwise, are unlikely to be limited to a single model. Consider data drift again: if any output of machine learning model A is used as an input to another machine learning model B, any degradation in accuracy in model A will propagate to model B and possibly be amplified in the process. As was the case with the data, we can detect such issues by using integration tests as quality gates to ensure that the inputs and the outputs of each model behave as expected. This is only possible if we track the dependencies between the models, for instance, by recording them as-code in the orchestrator configuration (Section 7.1.4) or by putting in place authentication and authorisation mechanisms to access models (say, with OAuth2 (ETF OAuth Working Group 2022)).

Therefore, we can say that technical debt at the model level arises mainly from feature and model entanglement: any issue that impacts one model’s inference capabilities will propagate to all the downstream models that depend on it, directly or indirectly, in what is called a correction cascade (Section 9.1.2). Entanglement between features, between models, and between features and models is unavoidable in practical applications: “changing anything changes everything” (Sculley et al. 2014). Features are rarely completely independent of each other, and black-box models (Section 9.2.2) like deep neural networks deliberately “entangle them” in ways that are difficult to understand. Models are also entangled with each other because they consume each other’s outputs (Section 9.1.2). This complex interplay unfortunately means that it can be difficult to find the root causes of the issues we are troubleshooting even when we observe tell-tale signs that something is wrong (Section 9.3).

On top of that, models are entangled with the real world: for instance, if the suggestions made by the model that drives a recommender system change, the behaviour of the system’s users will change in response. This creates a feedback loop because the users consume the model’s outputs and at the same time provide the data the model is trained on. Whether this is desirable or not depends on the specific application and on whether this feedback loop has a positive or negative effect: uncontrolled direct feedback loops can lead to an amplification of bias while artificially improving the model’s accuracy. Microsoft’s Tay chatbot (Hunt 2016) is a good case in point. Launched on Twitter in 2016 to “engage and entertain people through casual and playful conversation” while self-training from those conversations, it was shut down a few days later because every tweet it posted contained conspiracy theories or racist, inflammatory statements. (Maybe it maximised some abstract engagement metric in doing so?) Hidden feedback loops where machine learning models directly affect each other through exogenous events are also possible and harder to spot. Techniques such as reject inference (Crook and Banasik 2004) and contextual bandits (Dimakopoulou et al. 2018, 2019), collecting feedback from users and domain experts (Sections 5.3.4 and 5.3.5) and including additional features can help to break such loops by exploring new models and by suggesting whether the current ones should be retrained.

Finally, models may be entangled with each other when we take a pre-trained model and we fine-tune it for different tasks. This practice reduces computational requirements and speeds up model development: we buy a pre-trained model A for a general task (say, object detection) and then use tightly-focused data sets to specialise it into models B, C, etc. for specific tasks (say, detecting impurities in the semi-finished products of an industrial process). However, models B, C, etc. are likely to inherit similar failure modes from A, thus introducing coupling between models with no tracked dependencies and producing unexpected correction cascades in the machine learning pipeline. Furthermore, models B, C, etc. become more difficult to evolve independently because any bug we fix in model B should also be fixed in models A, C, etc. (or confirmed not to affect them) and the software tests for all models should be updated at the same time. Similarly, any enhancement that is meaningful for model B is likely to be meaningful for models A, C, etc. as well. We can manage these issues by using a configuration management platform, as we pointed out in Section 5.1, to track dependencies between models and between models and data, to version them and to enable systematic testing (Section 9.4.2).

5.2.3 At the Architecture (Design) Level

The architecture of a machine learning pipeline directs how data and models interact to achieve its goals: it is implemented as an orchestration system that schedules and coordinates various tasks such as data ingestion, data validation, feature engineering, model training and validation, model deployment on production systems, and serving. We will discuss them in detail in Section 5.3.

Machine learning pipelines are inherently complex systems with many moving parts, and they can easily hide architecture (design) debt. The key to keeping this type of technical debt in check is to give visibility into all aspects of their configuration as code using files in a human-readable data serialisation language like XML, YAML or JSON.12 These files should be under version control in a configuration management solution along with the data (Section 5.1) and the models (Section 5.2.2), and for similar reasons. Each change in design can then be expressed in those configuration files or using environment variables. Configuration files should be used for parameters, options and settings for which we need complete versioning across iterations, such as data set locations, training hyperparameters and model parameters. These files can also be linked to and supplement architecture documentation, which describes the pipeline using the more accessible ubiquitous language (Section 8.3). Environment variables should be used to store runtime configurations such as log-levels (Section 5.3.6), feature flags (Section 6.5) and the labels of target testing or production environments. Environment variables are also commonly used for secrets management, that is, to store credentials, certificates and other sensitive information. All modern software solutions to build machine learning pipelines provide mechanisms for configuring, overriding and exposing environment variables, including secrets. Only with a comprehensive formal description of the pipeline and of all its components we may be able to evolve and extend both over time without accidentally accruing architecture debt. Tracking and versioning the architecture along with the data and the models reduces the time spent on troubleshooting and debugging, and makes it possible to implement efficient deployment strategies (Section 7.2) and to roll back problematic models (Section 7.6). The alternative is to perform these operations manually, which is time-consuming and error prone: Knight Capital (.Seven 2014) proved that clearly to the world by burning $460 million in 45 minutes due to a botched manual deployment of their algorithmic trading software.

Unfortunately, we cannot control and version models from third-party libraries or remote systems as easily as those we train ourselves. Hence we are left to integrate them by wrapping their APIs with glue code to interface them with the rest of the machine learning pipeline. Glue code is a piece of ad hoc code, often in the form of a one-off script, that has no function other than to adapt software that would otherwise be incompatible. It is a common source of technical debt both at the model level (if shipped in the model) and at the architecture level (if used to bind together different modules in non-standard ways) where it creates what is known as the “pipeline jungle” anti-pattern (Bogner, Verdecchia, and Gerostathopoulos 2021).

Glue code is also commonly used to wrap libraries and remote APIs because it allows us to quickly expose them with new domain-specific names, interfaces and data structures (Section 8.2). While this practice may seem expedient, it can couple glue code tightly with what it is wrapping, causing it to break when the library or the remote API changes its public interface. We should only use glue code wrappers when we strictly need them, for example: to instrument a function for debugging purposes; to expose different versions or different features of the same library to different modules in the pipeline; or to integrate a legacy library or API that we would be otherwise unable to use.

5.2.4 At the Code Level

As for code debt, we should avoid mixing different versions of interpreters, programming languages and frameworks in the same machine learning pipeline. Unfortunately, this is a common issue for two reasons. Firstly, machine learning experts and data scientists often work in isolation, without a shared development environment. Secondly, microservices and similar architectures favour the use of multiple programming languages inside the same application in what they call polyglot programming. While it is often the case that different programming languages are better suited to different parts of a pipeline (Section 6.1), having too much variety can lead to organisational anti-patterns like an unbalanced distribution of skills and skill levels (say, there is only one developer with expertise in a key framework) and inadequate knowledge transfer (because there are too many technologies to keep track of). From a practical standpoint, a good compromise is to build any new machine learning pipeline from a small, up-to-date set of technologies and to involve all developers when incorporating new ones. The latter should be done sparingly: resume-driven development rarely ends well.

A related problem is that of vendoring software libraries, that is, including the source code of a specific version of a third-party software in our codebase instead of managing it as an external library through a package manager. Vendored libraries become untracked dependencies (Section 6.3), are often integrated using glue code, and are problematic to update because package managers and other automated tooling are unaware of their existence.

Another source of code debt is the amount of exploration and experimentation involved in creating machine learning models. It can easily produce dead experimental code paths, which are usually badly documented by comments (Section 8.1) and can lead to wasted effort as we try to achieve code coverage (Section 9.4.6). It can also limit the time we can spend on improving the quality of the code we produce from prototype to production level. Practices such as code review (Section 6.6) and constant refactoring (Section 6.7) can address both these issues, as we will discuss in the next chapter. They will also help in tackling low-quality code which, as a source of technical debt, significantly increases the number of bugs and the time required to fix them, slowing down development (Tornhill and Borg 2022).

5.3 Machine Learning Pipeline

Modern software development schools like Agile (Beck et al. 2001) and DevOps (Humble and Farley 2011) have pushed for the automation of testing, release management and deployment processes since the early 2000s, leading to the adoption of continuous integration / continuous delivery and deployment (CI/CD) solutions (Duvall, Matyas, and Glover 2007) to manage the software development life cycle. Continuous integration is the practice of developing code by committing small changes frequently to a version control repository. Each change is validated by an automated software testing solution, manually reviewed, and then integrated into the mainline branch the production builds are created from. As a result, the mainline branch is always in a working state and changes to the code are immediately visible to all developers. (More on that in Chapter 6.) Continuous delivery and continuous deployment focus on being able to release a working version of the software at any time and to deploy it on production systems. (More on that in Chapter 7.) In both cases, the emphasis is on using automated processes, versioning, configuration management, software testing and code review to enable an effortless, fast and reliable software development life cycle.

Life cycle of a machine learning pipeline.

Figure 5.2: Life cycle of a machine learning pipeline.

Nowadays, we have many integrated CI/CD solutions to build machine learning pipelines (called “MLOps”). However, a complete understanding of how a pipeline works becomes crucial when its development evolves from a simple proof of concept running on some developer’s local environment into a larger piece of software managed by a team and running on multiple systems. (Most real-world pipelines are complex enough to require a team to manage them.) At first, we explore some sample data and we try different models to gauge their performance, spending little to no time on software tests. Developing a pipeline then becomes the iterative and increasingly complex process shown in Figure 5.2: feeding new data from the ingestion phase to existing models for validating, monitoring and troubleshooting them; generating new models as the data change; deploying models and serving them continuously to downstream models or to the application or service that users will access. This is what we call a machine learning pipeline: the codification of these steps into independent, reusable, modular parts that can be pipelined together to orchestrate the flow of data into, and outputs from, machine learning models.13 MLOps practices standardise and automate how a pipeline is developed, giving us all the advantages that CI/CD brought to traditional software engineering, and builds on the same foundations: effective use of versioning, configuration management, automated testing, code review and automated deployments. Continuous integration, in addition to the testing and validation of code, now covers the testing and validation of the data and the models. Continuous delivery and continuous deployment expand to the production and deployment of the entire machine learning pipeline, again including the models. This extended definition of CI/CD allows us to focus on the development, testing and validation of the machine learning models, replacing homegrown solutions based on glue code with systematic solutions based on industry standards.

Figure 5.2 takes the software development life-cycle representation from Figure 1.2 and puts it into context. It shows the key logical steps of reproducible machine learning: what we should take care of to build a solid and maintainable pipeline. Some boxes represent development stages, some are actual pieces of software that will become modules in our pipeline, others are both. Broadly speaking, we can group the modules in a pipeline into four stages: data ingestion and preparation; model training, evaluation and validation; model deployment and serving; and monitoring, logging and reporting. How the functionality provided by each stage is split into modules is something that we can decide when we define the scope of the pipeline; we can then produce a baseline implementation to develop an understanding of its size and structure. However, well-established design principles from software engineering apply (Ousterhout 2018; Thomas and Hunt 2019). Each module should do one thing and do it completely (the “Single Responsibility Principle”), encapsulating as much complexity as possible and abstracting it behind a simple interface (a “deep module”). Thus, we can keep the complexity of the pipeline in check by avoiding change amplification (making a simple change requires modifying code many different locations) and by reducing cognitive load (how much does a developer need to know in order to successfully make the change) as well as unknown unknowns (which parts of the code should be touched is not obvious). Simple interfaces are less likely to change: they also reduce coupling between the modules if we limit the number of dependencies and avoid common anti-patterns such as implicit constraints (say, functions should be called in a specific order) and pass-through variables containing all kinds of unrelated information (say, the whole global state in a context object). Simple interfaces should also reflect domain knowledge by exposing methods and data structures with domain meaning, with names taken from the ubiquitous language (Chapter 8) and with default settings that make common cases simple to implement. This approach is likely to result in a pipeline architecture patterned after the workflow of domain experts, which allows them to help validate models and inference outputs in a “human-in-the-loop” setup (Wu et al. 2022; Xin et al. 2018). Furthermore, a modular pipeline can be easily managed by an orchestrator which can deploy the modules (Chapter 7), allocate them to systems with the appropriate hardware resources (Chapter 2) and control their execution.

5.3.1 Project Scoping

Starting from the top of Figure 5.2, the first step in building a machine learning pipeline is to understand the problem it should solve, what data it can use to do so, what outputs it should produce, and who its end users will be. To clarify these points, we should first identify who will be involved in developing the pipeline or will interact with it (the “stakeholders”): a combination of software developers, machine learning experts, domain experts and users. Together they will have all the information necessary to define the scope of the pipeline.

The process of scoping a machine learning pipeline and the underlying systems (Chapter 2) involves the following steps:

  1. Identifying the problem we want to solve: the stakeholders should work together to explicitly define the problem that the pipeline should solve and to evaluate its impact. Domain experts should have a concrete business or academic need to address and, together with the other stakeholders, they should decide whether the problem is worth solving and whether solving it will be valuable to enough people. This process is much smoother if the domain experts have some familiarity with the classes of problems that can be effectively tackled with machine learning.

  2. Identifying the targets we want to optimise for: the stakeholders should decide what it means to have solved the problem successfully. To this end, the domain experts should set measurable domain metrics with achievable threshold values to define “success”. These metrics should be:

    • comparable across different data, models and technical solutions to make it possible to contrast different pipeline implementations;
    • easy to understand and to interpret;
    • simple enough that they can be collected in real-time for logging and monitoring (Section 5.3.6);
    • actionable.

  1. Identifying what data we need: data are a critical component of a machine learning pipeline because they determine its performance (Section 5.1). Therefore, it is essential to identify all the data sources we want to use, who owns them, and the technical details of how the data are stored (files, databases or data lakes) and structured (data schema). This allows us to track data provenance and reduce technical debt (Section 5.2.1). In particular, we should be wary about data sources that provide overlapping information because they introduce hidden dependencies in the pipeline. They can easily be inconsistent because of differences in their schemas (say, the same variable is scaled or discretised in different ways) and, even if they are consistent, they can diverge over time (say, one data source changes schema and the others do not). A common case is that of partially pre-processed data, which should always be reconciled with the raw data they originate from and stored in the same versioned repository. In addition, we should collect data following the best practices accumulated in decades of survey sampling (Lohr 2021; Groves et al. 2009) and experimental design (Montgomery 20AD) to make sure that the data we collect to train the machine learning models (Section 5.3.4) are representative of the data the models will perform inference on (Section 5.3.5). Sampling bias can have unpredictable effects on the performance of the pipeline.

  2. Analysis: we should assess how much data we can collect and what variable types they will contain. With this information, we can start evaluating different models based on their sample size requirements, their probabilistic assumptions and the inference types they support (prediction, classification, etc.). As a general rule, it is always preferable to start with simpler models because they enable a fast feedback loop: if simple models cannot achieve our targets, we can move to more complex models and use the simpler ones as baselines. In addition, we should take into consideration:

    • The robustness of the model against the noise in the data, against model misspecification and adversarial attacks.

    • Interpretability and explainability, that is, how well we can understand the behaviour and the outputs of the models. Some models are inherently interpretable either because of their simple structure (say, regression models) or because of their construction (say, Bayesian networks (Scutari and Denis 2021)). For others (say, deep neural networks), we can introduce auxiliary models to provide post hoc explanations: some of them are application-agnostic (Linardatos, Papastefanopoulos, and Kotsiantis 2021) while others are specific to natural language processing (Li et al. 2016) or computer vision (Simonyan, Vedaldi, and Zisserman 2014).

    • The fairness of model outputs, which should not induce the machine learning pipeline to discriminate against individuals or groups based on sensitive attributes such as gender, race or age. While there is much literature on this topic (Mehrabi et al. 2021), there is no consensus on how fairness should be measured. What there is consensus on is that machine learning models can easily incorporate the biases present in the data they are trained from. Therefore, we should consider carefully how the data are collected and we should constrain models to limit or disregard the discriminating effect of known sensitive attributes. Failures to do so have often ended in the news: Amazon’s sexist recruitment tool (BBC 2018), Facebook image recognition labelling black men as primates (BBC 2021a) and Twitter’s racist preview cropping (BBC 2021b) are just a few examples.

    • Privacy and security concerns for sensitive data (Papernot et al. 2018). Machine learning models excel at extracting useful information from data, but at the same time, they should protect privacy by not disclosing personally identifiable information. How to achieve that is an open problem, with research investigating approaches like differential privacy (Gong et al. 2020), defences against adversarial attacks and data re-identification (Narayanan and Shmatikov 2008), and distributed learning implementations such as federated learning (Li et al. 2021) and edge computing (Khan et al. 2019) (Section 2.3).

A machine learning pipeline typically spans several data sources and several models: as a result, we will iterate over these steps a few times depending on the nature of the project and of the organisation undertaking it. In the end, we will have the information we need to compile a mission statement document (Section 8.4) and to sketch the layout of the architecture (Section 8.3) and of our software test suite (Section 9.4.1). The architecture is typically represented with a directed acyclic graph (DAG): see Figure 8.2 for an illustrative example. Each node will correspond to one of the modules in the pipeline, with incoming and outgoing arcs showing its inputs and outputs, respectively. The DAG therefore maps the paths of execution of the pipeline and the flow of data and information from data ingestion to training, inference and reporting. The DAG may be quite large for particularly complex pipelines: splitting it into smaller DAGs corresponding to different sections of the pipeline and working with them independently may be more convenient.

5.3.2 Producing a Baseline Implementation

Data validation, model development, tuning, training and validation are initially explored by individual developers and machine learning experts on local hardware, if suitable hardware is available. After experimentation, they will eventually produce a minimal, working prototype of some part of the pipeline. This is often called a baseline implementation or proof of concept, and it will only involve the smallest amount of code that allows us to check whether we can achieve our targets.

This initial exploration of the problem does not typically involve all the CI/CD development workflows discussed above and in Chapter 6: at this stage, the code and the models are too volatile. However, developers and machine learning experts should at least agree on a common, unified development environment (software dependencies management, build processes and configurations). This environment should be buildable in a reproducible and reliable way, which requires configuration management, and it should be as close as possible to our target production environment. For convenience, the development environment should be modular in the same way as the pipeline, so that we can run only the modules we are working on: it is typically impossible to run the whole pipeline on a developer workstation.

After checking that our proof of concept achieves all its targets, we then:

  1. Construct a suite of software tests (Section 9.4.2) and push both to our version control repository to start taking advantage of continuous integration. We can then transform the proof of concept into production-quality code by gradually refactoring (Section 6.7) and documenting it (Chapter 8) with the help of code review (Section 6.6).

  2. Improve scalability. A proof of concept is typically built using a small fraction of the available data, so we must ensure that its computational complexity (Chapter 4) is small enough to make learning and inference feasible in production when all data are used. Time complexity is important to allow for timely model retraining and for inference under latency constraints; space complexity must fit the machine learning systems (Chapter 2) we have available. If our development system is similar to the production systems, we can expect computational complexity to translate into practical performance in similar ways and predict the latter reliably.

5.3.3 Data Ingestion and Preparation

After scoping the pipeline and producing a baseline implementation of its parts, we can start designing and implementing its modules in a more structured way. A machine learning pipeline is the formalisation of a data processing workflow. Therefore, the first part of the pipeline will comprise one or more data ingestion modules where we collect data from various sources such as relational databases, legacy OLTP/OLAP systems and modern in-house or cloud data lakes. These modules vary in nature depending on the machine learning systems the pipeline will run on: their design will be heavily influenced by factors such as data locality (Sections 2.2 and 2.3), data provenance (Section 5.2.1), the availability of different types of storage (Section 2.1.2) and compliance with privacy frameworks like HIPAA and FCRA in the United Stated or GDPR in Europe (Section 5.3.1).

Data ingestion is followed by data preparation. Preparing and cleaning the data is a hard but crucial step involving data scientists, domain experts and machine learning experts (Kenett and Redman 2019). Modules for data preparation build on the exploratory analysis of the data used to produce the baseline implementation of the models, which is often limited to a high-level analysis of summary statistics, graphical visualisations and some basic feature selection. Their purpose is to clean and improve the quality of the data in the most automatic and reproducible way possible, making subsequent stages of the pipeline more reliable. In addition to validating the types, the acceptable values and the statistical distribution of each feature, data preparation modules should address the issues discussed in Section 9.1. They can also automate both feature selection and feature engineering (that is, the transformation of existing features into new ones that are better suited to model training or that are more meaningful in domain terms). Current software solutions for data and machine learning pipelines handle these tasks in a flexible way by taking as configuration arguments a processing function and a validation function that checks the properties of the now-clean data. The former may, for example, remove outliers, impute missing data and sort labels and features; the latter serves as a quality gate (Section 5.2.1) and as the kernel of a property-based software test (Section 9.4.2).

Finally, the data are split into multiple sets for later use as training, validation and test sets. (Making sure to avoid data leakage, see Section 9.3.) Each data set is tagged with information about its origin and with the version of the code that was used to extract and clean it, to track data provenance. These tags become part of our configuration management, and the data is stored as an artefact under versioning for later use.

5.3.4 Model Training, Evaluation and Validation

After ingestion and preparation, a machine learning pipeline passes the data either to model training modules or to inference modules (which we will discuss in Section 5.3.5). The trained models are then evaluated (on their statistical performance) and validated (in domain terms) using software tests and human expert judgement to ensure they are efficient, reproducible and scalable. Only models that perform sufficiently well in both statistical and domain terms will be considered suitable for deployment and serving.

Training a machine learning model consists in identifying an optimal instance in some model class (neural networks, random forests, etc.) by iteratively applying a combination of feature engineering, hyperparameter tuning and parameter optimisation. This is what the “learning” in “machine learning” refers to: a computer system is trained to learn a working model of some piece of the real world from the information contained in the data. The probabilistic techniques used for this purpose are specific to each model class and are beyond the scope of this book: see Kuhn and Johnson (M. Kuhn and Johnson 2013) for an approachable treatment of this topic. Training is a computationally demanding task, especially in the case of deep learning. The role of the pipeline is to schedule the training workload on compute systems with the appropriate hardware capabilities (as discussed in Section 2.4) and to monitor its progress. It should also simplify the parallel training of models with predefined, regular patterns of hyperparameters; and it should automate software tests implementing property-based testing of the model’s probabilistic properties (Section 9.4.2).

Training can take quite different forms depending on the nature of the data (Section 9.4.3). In static learning, the model is trained from scratch on cold (offline) data selected to be representative of the data currently observed in production. Its statistical performance is then evaluated against either a separate set of cold data or a small stream of production data. In either case, the data should be labelled or validated by domain experts to address the issues discussed in Section 5.2.1 and to maximise model quality. In dynamic learning, the model is continuously trained and evaluated on a live stream of (online) production data collected in real time. This requires fine-grained monitoring to be in place (Section 5.3.6). If data drift is gradual, we may prevent the model from going stale by fine-tuning it (Gama et al. 2014). If, on the other hand, data drift is sudden, it may be preferable to retrain the model from scratch with a batch of recent data.

Model evaluation modules check whether the predictive accuracy of the model the pipeline just trained is better in statistical terms than that of the corresponding model currently in production. To assess both simultaneously, we can perform a canary deployment: running the current and the new model in parallel on the same data to compare them directly. (More on this in Chapter 7.) In the case of streaming data, it is standard practice to use A/B testing (Amazon 2021; Zheng 2015) for this purpose, assigning new data points at random to either model. At the same time, we can check whether the new model is preferable to the current one in domain terms using the metrics we decided to optimise for (Section 5.3.1). We call this model validation, in contrast with the evaluation of the model in purely statistical terms. The two may be related because models with poor statistical properties will typically not encode the domain well enough for practical use. However, models with good statistical properties are not always of practical use either: in particular when the loss function the model is trained to minimise is too different from that implied by how costly prediction errors are in business or domain terms. In general, it is better to choose well-matched domain metrics and statistical accuracy measures for consistency. Unlike model evaluation, which can be automated to a large extent using software tests and continuous integration, model validation should involve domain experts. Even if we practise domain-driven development (Evans 2003) and involve them in the design of the pipeline, in implementing it (Chapter 6) and in documenting it (Chapter 8), there will always be some domain knowledge or intuition that they were not able to convey to developers and machine learning experts. As unscientific as it may sound, there is knowledge that is essentially impossible to put into numbers. Therefore, there will be issues we cannot write tests for, but that experts can “eyeball” and flag in model outputs because “they look wrong” and “do not quite make sense.” This approach is known as “human-in-the-loop” in the literature, and it is known to improve the quality of machine learning across tasks and application fields (Wu et al. 2022; Xin et al. 2018).

When a model is finally found to perform well in both statistical and domain terms, the pipeline should trigger a CI/CD process to generate an artefact containing the model and all the relevant information from the training process. An artefact can be, from simple to complex:

  1. A (usually binary) file in a standardised format that will be stored and versioned in a general-purpose artefact registry. The format can be either model-independent, like ONNX (ONNX 2021), or specific to the machine learning framework used for training.
  2. A (usually Docker (Docker 2022a)) container that embeds the model and wraps it with application code that provides APIs for inference, health checking and monitoring. The container is then stored and versioned in a container registry.
  3. An annotated file uploaded to a model registry that provides experiment tracking, model serving, monitoring and comparison between models in addition to versioning.

Platforms like GitHub and GitLab integrate both a general-purpose artefact registry (GitHub 2022b; GitLab 2022a) and a container registry (GitHub 2022c; GitLab 2022b), as does Nexus (Sonatype 2022). MLOps platforms like TensorFlow Extended (TFX) (TensorFlow 2021b) implement experiment tracking and other machine-learning-specific features. We will return to this topic in Section 7.1.

Regardless of their form, artefacts should be immutable: they cannot be altered once generated so they can be used as the single source of truth for the model. Data artefacts (Section 5.3.3), code (Section 6.5) and often other software artefacts are also stored as immutable artefacts and versioned. When their versions are linked, we have a complete configuration management solution that allows for reproducible builds of any development, testing or production environment that has ever been used in the pipeline.

5.3.5 Deployment, Serving and Inference

Not all the artefacts we produce will be deployed immediately, or at all: continuous delivery only ensures that we are always ready to deploy our latest models. In academia, we cannot make any change to a pipeline halfway through a set of experiments without potentially introducing confounding in the results. In business, we may have service-level agreements with our customers that make it risky to deploy new models without a compelling reason to do so. Artefacts may also be found to be unsuitable for deployment for security reasons: for instance, we may find out that a container contains vulnerable dependencies or is misconfigured (Section 7.1.4).

Model deployment is not implemented as a module: rather, it is the part of the pipeline orchestration that enables models to be deployed to a target environment. Models deployed in production will be served so that users, applications or other models can access their inference capabilities. Models deployed to test environments will be evaluated by software tests and expert judgement, and those deployed to development environments can be used for troubleshooting bugs or further investigation of the data.

How a machine learning model is deployed depends on how it has been packaged into an artefact and on how it will be used. File artefacts can be either embedded in a software library that exposes inference methods locally or served “as-a-service” from a model registry using suitable remote APIs and protocols (such as RESTful or, when we need low latency, gRPC (Ganiev et al. 2021)). Container artefacts can be deployed by all orchestration platforms in common use, which provide built-in monitoring and logging of hardware and software metrics (load, memory and I/O use) as well as troubleshooting facilities. Despite being intrinsically more complex, container artefacts are easier to deploy because they are ephemeral and highly portable, and because we can manage as-a-code both their runtime dependencies and configuration. We will develop this topic in detail in Sections 7.1.4 and 7.2 using Dockerfiles as a reference.

5.3.6 Monitoring, Logging and Reporting

Monitoring modules collect the metrics we identified in the scoping phase (Section 5.3.1) to track at all times whether the pipeline achieves the required statistical and domain performance levels. The metrics should describe both the pipeline as a whole and individual modules to allow us to pinpoint the source of any issue we may have to troubleshoot. In particular:

  • Data ingestion and preparation modules (Section 5.3.3): we should monitor the same data metrics we check with property-based software tests to guard against data drift and data degradation.
  • Training modules (Section 5.3.4): we should monitor the same metrics we use for model validation and evaluation consistently across all models in the pipeline to separate issues with individual models from issues arising from the data. Especially when using online data.
  • Serving and inference modules (Section 5.3.5): we should monitor the same metrics we monitor during training to ensure that performance has not degraded over time (the so-called “training-serving skew”). And we should do that for all inference requests (possibly in small batches) so that we can guarantee that outputs are always in line with our targets. This is crucial to enable human-in-the-loop validation by domain experts for black-box models whose failure modes are mostly unknown and difficult to test.

The coverage of monitoring facilities is important for the same reason why test coverage is important: both are tasked to identify a broad range of issues with the data (Section 9.1), with the models (Section 9.2) and with the pipeline (Section 9.2.4) with enough precision to allow for root-cause analyses. Software tests perform this function at development and deployment time; monitoring does it at runtime.

In practice, we can implement monitoring with a client-server software such as Prometheus (Prometheus Authors and The Linux Foundation 2022). Each module in the pipeline produces all relevant metrics internally, tags them to track provenance (which module, and which instance of the module if we have multiple copies running in parallel) and makes them available in a structured format through the client interface. Monitoring modules then provide the corresponding server that pulls the metrics from all clients and saves them into an event store database. They will also filter the metrics, sanitise them, and run frequent checks for anomalies. If any is found, the monitoring modules can then trigger alerts and send failure reports to the appropriate people using, for instance, Alertmanager (which is part of Prometheus) or PagerDuty (PagerDuty 2022). If our pipeline is sufficiently automated, we may also trigger model retraining automatically at the same time. This is the only way to address anomalies in a timely manner and to provide guarantees on the quality of the outputs of the pipeline. Cross-referencing the information in the event store to that in our configuration management system is invaluable in comparing the performance of our current production environment against that of past (now unavailable) environments. The same metrics may also be useful for troubleshooting infrastructure issues, like excessive consumption of computing resources, memory and I/O, as well service issues that impact downstream services and models, like readiness (whether a specific API is ready to accept requests) and excessive inference latency (how long it takes for the API to respond).

Logging modules complement monitoring by recording relevant information about events that occur inside individual modules or within the pipeline orchestration, capturing exceptions and errors. Typically, at least part of a machine learning pipeline runs on remote systems: since we cannot access them directly, especially in the case of cloud instances (Section 2.3), we are limited in our ability to debug and troubleshoot issues. Logging makes this problem less severe by recording what each module is doing in a sequence of timestamped log messages, ranging from simple plain-text messages (as we may produce ourselves) to more structured JSON or binary objects (from frameworks or language interpreters). Each log message has a “level” that determines its severity and that allows us to control how much we want to log for each module: for instance, a set of labels like DEBUG, INFO, WARNING, ERROR, and CRITICAL. Each log message is also tagged with its provenance, which allows us to distinguish between:

  • system logs, which provide information on the load of the machine learning systems, the runtime environment and the versions of relevant dependencies;
  • training logs, which describe the model structure, how well it fits the data and the values of its parameters and hyperparameters for each training iteration;
  • inference logs, which describe inputs, outputs, accuracy and latency for each request and each API.

Therefore, logs provide a measure of observability when we otherwise would have none: all modules should implement logging as much as monitoring. However, the more messages we generate, the more resources logging requires: which poses practical limits on how much we can afford to log, especially on production systems. In development environments, we may just append log messages to a file. In production environments, we should aggregate log messages from the whole pipeline to a remote log collector instead of locally. Log collectors can normalise log messages, make them easy to browse and make it possible to correlate events happening in different modules.

Similar to monitoring modules, logging modules are implemented with a client-server software such as Fluentd (The Fluentd Project 2022) complemented by a search engine like Elasticsearch and a web frontend like Kibana (Elasticsearch 2022). The two software stacks have some apparent similarities: both have a remote server aggregating information from clients inside the modules. The underlying reason for this architecture is that we should locate the server on a system that is completely separate from those the machine learning pipeline runs on: when the latter crashes and burns, we need to be able to access the information stored by monitoring and logging servers to investigate what its last known status was and decide how to best restore it.

However, monitoring and logging have two key technical differences. Firstly, logging should support unstructured data, whereas monitoring only handles data in the form of {key, type, value} triplets. Logging gives observability from outside the code we wrote to implement a module, reporting information that we do not produce directly and whose format we cannot necessarily control. Monitoring gives observability from the inside: we incorporate the client component into our code and we give it access to its internal state. Hence the information we expose to the monitoring server is necessarily structured in various data types and data structures (Chapter 3). Secondly, logs are pushed from the clients to the servers as they are generated, whereas monitoring servers pull the metrics from the clients in the modules at regular intervals. Therefore, the databases used by the logging servers are general-purpose event stores, whereas those used for monitoring are optimised for time series data. The ability to access the internal state of all modules at regular intervals makes monitoring servers ideal for observing any gradual degradation in the machine learning pipeline.

Reporting modules implement graphical interfaces that display the information collected by the monitoring and logging modules. Building on best practices from data science (Kenett and Redman 2019), they provide web interfaces with intuitive, interactive dashboards that can be used by developers, machine learning experts and domain experts alike. Graphical displays in common use are:

  • Data ingestion and preparation modules (Section 5.3.3):
    • Plots of the empirical distribution both of individual features and of pairs of features against each other such as histograms, boxplots, heatmaps and pairwise scatterplots (for continuous features) or barplots and tileplots (for discrete features).
    • Plots of key summaries from minimal statistical models such as simple linear regressions to assess the magnitude and the sign of the relationships between features and to explore potential fairness issues.
  • Training modules (Section 5.3.4):
    • Plots of model performance over the course and at the end of the training process, like profile plots of the loss function against epochs for deep neural networks and heatmaps for confusion matrices produced by classification models.
    • Plots that help interpret the model behaviour, showing either its parameters or the outputs of explainability approaches like LIME (Ribeiro, Singh, and Guestrin 2016) and SHAP (Lundberg and Lee 2017).
    • For less computationally-intensive models, interactive dashboards that can trigger model training, with sliders to pick hyperparameters on the fly.
  • Serving and inference modules (Section 5.3.5):
    • Plots of the empirical distribution of input data against historical data, to detect data drift.
    • Time series plots of the accuracy measures used in model validation and the metrics used for model evaluation, to detect when models become stale.
    • Time series plots of latency and readiness.

All plots should also include confidence intervals to convey the likely range of values for each of the quantities they display, wherever it makes sense.

Domains like natural language processing and computer vision may require specialised graphical interfaces in addition to the above: for instance, visualising word relevance in natural language processing (Li et al. 2016) and pixel relevance in computer vision (Simonyan, Vedaldi, and Zisserman 2014) or splitting images into layers with semantic meaning (Ribeiro, Singh, and Guestrin 2016). Such interfaces can be very useful to involve domain experts in validating model training and the outputs from the inference modules. Instances that were not classified or predicted correctly can then be visually inspected, labelled and used to retrain the machine learning models.

References

Amazon. 2021. Dynamic A/B Testing for Machine Learning Models with Amazon SageMaker MLOps Projects. https://aws.amazon.com/blogs/machine-learning/dynamic-a-b-testing-for-machine-learning-models-with-amazon-sagemaker-mlops-projects/.

Arpteg, A., B. Brinne, L. Crnkovic-Friis, and J. Bosch. 2018. “Software Engineering Challenges of Deep Learning.” In Euromicro Conference on Software Engineering and Advanced Applications, 50–59. IEEE.

BBC. 2018. Amazon Scrapped “Sexist AI” Tool. https://www.bbc.com/news/technology-45809919.

BBC. 2021a. Facebook Apology as AI Labels Black Men “Primates”. https://www.bbc.com/news/technology-58462511.

BBC. 2021b. Twitter Finds Racial Bias in Image-Cropping AI. https://www.bbc.com/news/technology-57192898.

Beck, K., M. Beedle, A. Van Bennekum, A. Cockburn, W. Cunningham, M. Fowler, J. Grenning, et al. 2001. The Agile Manifesto. https://www.agilealliance.org/wp-content/uploads/2019/09/agile-manifesto-download-2019.pdf.

Bogner, J., R. Verdecchia, and I. Gerostathopoulos. 2021. “Characterizing Technical Debt and Antipatterns in AI-Based Systems: A Systematic Mapping Study.” In 2021 IEEE/ACM International Conference on Technical Debt (TechDebt), 64–73.

Cheney, J., L. Chiticariu, and W.-C. Tan. 209AD. “Provenance in Databases: Why, How and Where.” Foundations and Trends in Databases 1 (4): 379–474.

Crook, J., and J. Banasik. 2004. “Does Reject Inference Really Improve the Performance of Application Scoring Models?” Journal of Banking and Finance 28: 857–74.

Cunningham, W. 1992. “The WyCash Portfolio Management System.” In Addendum to the Proceedings of ACM Object-Oriented Programming, Systems, Languages & Applications Conference, 29–30.

Cunningham, W. 2011. Ward Explains the Debt Metaphor. https://wiki.c2.com/?WardExplainsDebtMetaphor.

Dimakopoulou, M., Z. Zhou, S. Athey, and G. Imbens. 2018. Estimation Considerations in Contextual Bandits. https://arxiv.org/abs/1711.07077.

Dimakopoulou, M., Z. Zhou, S. Athey, and G. Imbens. 2018. Estimation Considerations in Contextual Bandits. https://arxiv.org/abs/1711.07077.

2019. “Balanced Linear Contextual Bandits.” In Proceedings of the AAAI Conference on Artificial Intelligence, 3445–53.

Docker. 2022a. Docker. https://www.docker.com/.

Duvall, P. M., S. Matyas, and A. Glover. 2007. Continuous Integration: Improving Software Quality and Reducing Risk. Addison-Wesley.

Elasticsearch. 2022. Free and Open Search: The Creators of Elasticsearch, ELK & Kibana. https://www.elastic.co/.

ETF OAuth Working Group. 2022. OAuth 2.0. https://oauth.net/2/.

Evans, E. 2003. Domain-Driven Design: Tackling Complexity in the Heart of Software. Addison-Wesley.

Gama, J., I. Žliobaitè, A. Bifet, M. Pechenizkiy, and A. Bouchachia. 2014. “A Survey on Concept Drift Adaptation.” ACM Computing Surveys 46 (4): 44.

Ganiev, A., C. Chapin, A. Andrade, and C. Liu. 2021. “An Architecture for Accelerated Large-Scale Inference of Transformer-Based Language Models.” In Proceedings of the 2021 Conference of the North American Chapter of the Association for Computational Linguistics, 163–69.

GitLab. 2022a. GitLab Artifacts.

GitLab. 2022b. GitLab Container Registry. https://docs.gitlab.com/ee/user/packages/container_registry/.

Gong, M., Y. Xie, K. Pan, and K. Feng. 2020. “A Survey on Differentially Private Machine Learning.” IEEE Computational Intelligence Magazine 15 (2): 49–64.

Groves, R. M., F. J. Fowler, M. P. Couper, J. M. Lepkowski, E. Singer, and R. Tourangeau. 2009. Survey Methodology. Wiley.

Humble, J., and D. Farley. 2011. Continuous Delivery. Addison Wesley.

Hunt, E. 2016. Tay, Microsoft’s AI Chatbot, Gets a Crash Course in Racism from Twitter. https://www.theguardian.com/technology/2016/mar/24/tay-microsofts-ai-chatbot-gets-a-crash-course-in-racism-from-twitter.

Kenett, R. S., and T. C. Redman. 2019. The Real Work of Data Science. Wiley.

Khan, W. Z., E. Ahmed, S. Hakak, I. Yaqoob, and A. Ahmed. 2019. “Edge Computing: A Survey.” Future Generation Computer Systems 97: 219–35.

Kuhn, M., and K. Johnson. 2013. Applied Predictive Modeling. Springer.

Li, J., X. Chen, E. Hovy, and D. Jurafsky. 2016. “Visualizing and Understanding Neural Models in NLP.” In Proceedings of the 2016 Conference of the North American Chapter of the Association for Computational Linguistics: Human Language Technologies, 681–91. Association for Computational Linguistics.

Li, Q., Z. Wen, Z. Wu, S. Hu, N. Wang, Y. Li, X. Liu, and B. He. 2021. “A Survey on Federated Learning Systems: Vision, Hype and Reality for Data Privacy and Protection.” IEEE Transactions on Knowledge and Data Engineering Advance publication.

Linardatos, P., V. Papastefanopoulos, and S. Kotsiantis. 2021. “Explainable AI: A Review of Machine Learning Interpretability Methods.” Entropy 23 (1): 18.

Lohr, S. L. 2021. Sampling: Design and Analysis. 3rd ed. CRC Press.

Lundberg, S. M., and S.-I. Lee. 2017. “A Unified Approach to Interpreting Model Predictions.” In Advances in Neural Information Processing Systems (NIPS), 4765–74.

Mehrabi, N., F. Morstatter, N. Saxena, K. Lerman, and A. Galstyan. 2021. “A Survey on Bias and Fairness in Machine Learning.” ACM Computing Surveys 54 (6): 115.

Montgomery, D. C. 20AD. Design and Analysis of Experiments. 10th ed. Wiley.

Narayanan, A., and V. Shmatikov. 2008. “Robust De-Anonymization of Large Sparse Datasets.” In Proceedings of the IEEE Symposium on Security and Privacy, 111–25.

ONNX. 2021. Open Neural Network Exchange. https://github.com/onnx/onnx.

Ousterhout, J. 2018. A Philosophy of Software Design. Yaknyam Press.

PagerDuty. 2022. PagerDuty: Uptime Is Money. https://www.pagerduty.com/.

Papernot, N., P. McDaniel, A. Sinha, and M. P. Wellman. 2018. “SoK: Security and Privacy in Machine Learning.” In Proceedings of the IEEE European Symposium on Security and Privacy, 399–414.

Paszke, A., S. Gross, F. Massa, A. Lerer, J. Bradbury, G. Chanan, T. Killeen, et al. 2019. “PyTorch: An Imperative Style, High-Performance Deep Learning Library.” In Advances in Neural Information Processing Systems (Nips), 32:8026–37.

Prometheus Authors, and The Linux Foundation. 2022. Prometheus: Monitoring System and Time Series Databases. https://prometheus.io/.

Ribeiro, M. T., S. Singh, and C. Guestrin. 2016. “Why Should I Trust You? Explaining the Predictions of Any Classifier.” In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, 1135–44. ACM.

Scikit-learn Developers. 2022. Scikit-learn: Machine Learning in Python. https://scikit-learn.org/.

Sculley, D., G. Holt, D. Golovin, E. Davydov, T. Phillips, D. Ebner, V. Chaudhary, and M. Young. 2014. “Machine Learning: The High Interest Credit Card of Technical Debt.” In SE4ML: Software Engineering for Machine Learning (NIPS 2014 Workshop).

Scutari, M., and J.-B. Denis. 2021. Bayesian Networks with Examples in R. 2nd ed. Chapman & Hall.

.Seven, D. 2014. Knightmare: A DevOps Cautionary Tale. https://dougseven.com/2014/04/17/knightmare-a-devops-cautionary-tale/.

Sherman, E. 2022. What Zillow’s Failed Algorithm Means for the Future of Data Science. https://fortune.com/education/business/articles/2022/02/01/what-zillows-failed-algorithm-means-for-the-future-of-data-science/.

Simonyan, K., A. Vedaldi, and A. Zisserman. 2014. “Deep Inside Convolutional Networks: Visualising Image Classification Models and Saliency Maps.” In Proceedings of the 2nd International Conference on Learning Representations (ICLR), Workshop Track.

Sonatype. 2022. Nexus Repository Manager. https://www.sonatype.com/products/nexus-repository.

TensorFlow. 2021b. TensorFlow Extended (TFX). https://www.tensorflow.org/tfx/.

The Fluentd Project. 2022. Fluentd: Open Source Data Collector. https://www.fluentd.org/.

Thomas, D., and A. Hunt. 2019. The Pragmatic Programmer: Your Journey to Mastery. Anniversary. Addison-Wesley.

Tornhill, A., and M. Borg. 2022. “Code Red: The Business Impact of Code Quality: A Quantitative Study of 39 Proprietary Production Codebases.” In Proceedings of International Conference on Technical Debt, 1–10.

Wu, X., L. Xiao, Y. Sun, J. Zhang, T. Ma, and L. He. 2022. “A Survey of Human-in-the-Loop for Machine Learning.” Future Generation Computer Systems 135: 364–81.

Xin, D., L. Ma, J. Liu, S. Song, and A. Parameswaran. 2018. “Accelerating Human-in-the-Loop Machine Learning: Challenges and Opportunities.” In Proceedings of the Second Workshop on Data Management for End-to-End Machine Learning, 1–4.

Zheng, A. 2015. Evaluating Machine Learning Models. O’Reilly.


  1. By “traditional software”, we mean any software that is not related to analytics, data science or machine learning.↩︎

  2. The choice of the language is often dictated by the orchestration software. However, YAML is becoming a de facto standard because of its readability, portability and maturity.↩︎

  3. In software engineering, “pipeline” is used to mean the process of developing and delivering software: CI/CD is a pipeline. In this book, we use it to mean the software infrastructure to develop and put to use the machine learning models and, by extension, the process of building and operating it.↩︎