Skip to content
BY-NC-ND 4.0 license Open Access Published by De Gruyter May 20, 2017

Surgical process modeling

  • Thomas Neumuth EMAIL logo

Abstract

Due to the rapidly evolving medical, technological, and technical possibilities, surgical procedures are becoming more and more complex. On the one hand, this offers an increasing number of advantages for patients, such as enhanced patient safety, minimal invasive interventions, and less medical malpractices. On the other hand, it also heightens pressure on surgeons and other clinical staff and has brought about a new policy in hospitals, which must rely on a great number of economic, social, psychological, qualitative, practical, and technological resources. As a result, medical disciplines, such as surgery, are slowly merging with technical disciplines. However, this synergy is not yet fully matured. The current information and communication technology in hospitals cannot manage the clinical and operational sequence adequately. The consequences are breaches in the surgical workflow, extensions in procedure times, and media disruptions. Furthermore, the data accrued in operating rooms (ORs) by surgeons and systems are not sufficiently implemented. A flood of information, “big data”, is available from information systems. That might be deployed in the context of Medicine 4.0 to facilitate the surgical treatment. However, it is unused due to infrastructure breaches or communication errors. Surgical process models (SPMs) alleviate these problems. They can be defined as simplified, formal, or semiformal representations of a network of surgery-related activities, reflecting a predefined subset of interest. They can employ different means of generation, languages, and data acquisition strategies. They can represent surgical interventions with high resolution, offering qualifiable and quantifiable information on the course of the intervention on the level of single, minute, surgical work-steps. The basic idea is to gather information concerning the surgical intervention and its activities, such as performance time, surgical instrument used, trajectories, movements, or intervention phases. These data can be gathered by means of workflow recordings. These recordings are abstracted to represent an individual surgical process as a model and are an essential requirement to enable Medicine 4.0 in the OR. Further abstraction can be generated by merging individual process models to form generic SPMs to increase the validity for a larger number of patients. Furthermore, these models can be applied in a wide variety of use-cases. In this regard, the term “modeling” can be used to support either one or more of the following tasks: “to describe”, “to understand”, “to explain”, to optimize”, “to learn”, “to teach”, or “to automate”. Possible use-cases are requirements analyses, evaluating surgical assist systems, generating surgeon-specific training-recommendation, creating workflow management systems for ORs, and comparing different surgical strategies. The presented chapter will give an introduction into this challenging topic, presenting different methods to generate SPMs from the workflow in the OR, as well as various use-cases, and state-of-the-art research in this field. Although many examples in the article are given according to SPMs that were computed based on observations, the same approaches can be easily applied to SPMs that were measured automatically and mined from big data.

Surgical process models (SPMs) – motivation, definition, and delimitation

Worldwide, technological developments are gaining momentum and infiltrate more and more areas of everyday life. New information technologies, such as the Internet of Things or Industry 4.0, new miniature technologies, new energetic or storage systems, or the fact that intricate technology can be mass produced and thus is easily available around the globe at reasonable prices, transform our world profoundly. For the benefit of the patients and the promotion of human health, surgery and other medical disciplines are slowly merging with emerging technical disciplines, becoming more and more complex hybrids.

Overall, this development must be regarded as mostly positive, as it offers increasingly efficient treatment for patients, including enhanced patient safety and the treating possibilities of conditions that were regarded as untreatable before, minimally invasive interventions, and the prevention of errors. However, it also heightens the pressure under which surgeons and other medical professionals must operate. This situation also has engendered a new policy in hospitals, which must rely on a wide variety of factors, including economic, social, psychological, qualitative, practical, and technological factors.

In congruence with the quickly evolving medical, technological, and technical possibilities, surgical procedures are becoming ever more sophisticated [1]. This offers increasing advantages for patients, such as enhanced patient safety, wound care, and trauma management, as well as a reduction of scarring and of highly invasive therapies. However, the growing surgical sophistication also entails other developments.

Surgical procedures are becoming more and more complex. The reasons for this growing complexity are manifold. First, the growth of surgical knowledge, in general, entails a heightened number of possible surgical procedures, methods, and techniques that the surgeons must learn, remember, and practice. Second, the altered policy of medicine, in general – in accordance with the growing patients’ demands – has brought about a change of the overall objective of surgical interventions’ need to be minimal invasive, cost-efficient, and as easy on the patients as possible. The new policy therefore relies on more resources, such as economic [2], [3], [4], social, psychological, qualitative, practical, and technological aspects, enhancing patient safety and eliminating medical malpractices [5].

The technological factor is the one that has possibly had the greatest impact on modern surgery. Medical disciplines, especially surgery, are slowly merging with technical disciplines. The number of technical and technological appliances in the operating room (OR) is steadily on the rise, requiring the surgeons to adopt – in addition to their medical formation – a more technical background and to gather knowledge and experience regarding the best possible application of modern technology [6], [7] for the sake of the best possible outcome for the patients and to adhere to current best-practice as well as to economic aspects.

Although the OR, in general, is the most expensive hospital unit with regard to patient treatment [8], [9], the current information and communication technology (ICT) in hospitals is not able to support the clinical and operational sequence adequately [10], [11], [12], [13], thus creating breaches in the surgical workflow, prolongations in routines, additional staff or room idle times, and media disruptions. The growing amount of information gathered in the OR by means of technology, the big data, and the knowledge concerning surgical “modi operandi” that might support Medicine 4.0 are not sufficiently inspected, selected or discarded, interpreted, and employed to the advantage of the intervention at hand or to that of the overall management of all ORs and the hospital or the patients in general. Thus, it becomes clear that a more consistent, process-oriented support of the ICT is needed. This support can be founded on a basic and comprehensive describability of surgery and of surgical interventions.

The problem, however, is that surgical interventions entail a vast amount of variability and complexity. Therefore, conventional modeling approaches alone, such as the top-down modeling approach [14], [15], a method based on the experience of the process modelers and/or on interviews with experts of the domain in question [16], [17], abstracting the knowledge from its upper levels down to the smaller details, have been proven unfit for the task at hand. The main drawbacks of these approaches are their time-consuming, expensive, and subjective nature and the insufficient resolution, longevity, and quantifiability of the results. Another problem is that usually surgical processes are reviewed only from a single perspective, for instance, as a business process or as a workflow system. This can result in the exclusion of aspects that are important for the overall process.

Therefore, more adequate methods are needed for the acquisition and generalization of process models for surgical interventions and procedures, and more suitable possibilities for the employment of the models thus attained need to be devised to the advantage of surgeons, other hospital staff, the administration, and – last but not least – the patients. The development of advanced and validated methods for defining, gathering, saving, and documenting surgical processes out of the big data in the OR is necessary to improve the documentation and evaluation of surgical tasks. These innovative methods would make an identification of objective topics that influence the progression of surgical procedures possible and thus an assessment of the significance of emerging or existing technologies, pioneering surgical assist systems (SAS), or surgical techniques in general [18].

To summarize the argumentation so far, it can be said that accurate SPMs are very much needed to enhance modern surgical performance and to support Medicine 4.0 in the OR. Such models could help in the implementation of new surgical techniques, of evaluating and ameliorating existing techniques, and to organize, optimize, and manage the surgical process as well as adherent processes, such as OR scheduling or cost-planning. The problem, though, is that the required data are difficult to obtain and to structure. New approaches are needed to be able to compose models of surgical procedures that can be generalized and computed by information systems, can preserve patient specificity, and can be employed in a meaningful and customizable way. To achieve this aim, it is important to identify the clinical and technical motivation for such models as well as possible limitations. Furthermore, some definitions and a delimitation of notions such as surgical process, SPMs, and other connected concepts are needed.

Clinical motivation

It is possible to appraise certain facets of a surgical intervention objectively. Examples for such objective aspects are the duration of single work-steps or whole intervention phases or whether or not one approach is more effective than another approach. However, the overall evaluation of surgical practice is very intricate, complex, and highly subjective. There is no method available or described in the appertaining literature that allows for an objective and reliable quantification of surgical practice from a process point of view.

Nevertheless, such an appraisal is more important than ever for various reasons. One of these reasons is the context of the reduction of working hours for medical residents according to the changed working time directives of European law, which might induce an extra pressure on the surgeon to perform faster and to complete more interventions in a single working day. Furthermore, the number of medical – and foremost surgical – students is on the rise. This aggravates the already existing problem with providing enough practical training for each of them.

However, the problem expands to other areas as well. The current situation is such that increasing numbers of novel and innovative technologies become available. The implementation of these technologies into the medical routine with the aim of providing the best possible treatment for patients sometimes competes against ethical or financial concerns, especially when these technologies have not yet been subjected to a multicenter cost-benefit-ratio investigation. With regard to ethical concerns, one example would be a new technology that becomes available that would ameliorate the overall outcome of an intervention but that cannot be financed by reimbursements. In this case, the management of a hospital would have to choose between financial loss and ethical behavior. Effectively, it has to be stated that the overall standard of surgical training faces some serious issues, such as heightened workloads for surgeons, a need for more efficient training structures, increasing complexity of medical technological systems, increasing economic pressure, and a decreasing amount of time-per-patient.

Today, there is an urgent necessity to record, model, and analyze surgical practice by means of computers. This recording technology has to be objective and unbiased and needs to function robustly and unfalteringly. Furthermore, its implementation into a regular OR has to be well thought-through and practical so as not to take up valuable operating space and not to place a financial burden on the hospital or organization employing it. Furthermore, the use of the technology needs to be effortless so as not to disrupt daily clinical practice and heighten the pressure on surgeons and other OR staff. In addition, these systems need to be adjustable and adaptable to different surgical procedures.

The data thus collected might be used to describe the single surgical work-steps and the whole surgical procedures, with the aim of modeling the overall intervention, including all possible or all necessary information. In this regard, the term “modeling” can be used to support one or more of the following tasks: “to describe”, “to understand”, “to explain”, to optimize”, “to learn”, “to teach”, or “to automate”. The issuing models can be used to compare different types of interventions, such as approaches using different surgical tools or instruments, to compare single surgeons or whole OR teams, or to evaluate and compare different surgical procedures.

Technical motivation

From a technical point of view, SPMs can be used for different aims. One of these aims is requirements analyses and product specification. The advantage for engineers is the possible graphical representation of the results instead of a representation as text. Furthermore, such data are quantifiable, thus allowing for a derivation of technical or clinical constraints as, for instance, average durations of performance steps, total number of performance steps, or number of iteration cycles. Additionally, benchmark studies can be performed to compare surgical strategies, substitutive technologies, or approaches or to assess the skill level of resident surgeons in comparison to senior surgeons. Product comparisons are another use-case for SPMs. By means of these comparisons, for instance, the invasiveness of one surgical product can be evaluated in contrast to another product. In addition, the data can be used for risk and quality management. Furthermore, the SPMs can be used as input for workflow management systems so that the infrastructure and technical systems in the OR can be automatically triggered or controlled according to the actual surgical situation. These and other technical merits will be further described in “Model application examples”.

Terms and definitions

To define an SPM, various other notions have to be defined first, such as medicine, surgery, process, and model. However, the next paragraphs will only give a very brief introduction into this topic and not a comprehensive account, as this is not within the scope of this paper.

“Medicine” is the “science or practice of the diagnosis, treatment, or prevention of disease” [19]. As a branch of medicine, the overall goal of “surgery” is the same. To achieve this goal, surgery is using special means. Thus, surgery is a branch of medicine treating diseases and conditions in humans or animals using manual or instrumental procedures to diagnose, cure, or alleviate a pathological condition [19], [20], [21]. The term Medicine 4.0 is adapted from the definition of Industry 4.0 [22]. It emphasizes strategies and technologies for patient-specific individualization of treatment, hybridization of medical treatment and accompanied services, and the increasing involvement of stakeholders of the surgical treatment, such as hospital administration.

A “work-step” or activity” can be defined as “a description of a piece of work that forms one logical step within a process” [23]. In consequence, a “process” can be defined as a “network of activities and their relationships, criteria to indicate the start and termination of the process, and the information about the individual activities, such as participants” [23]. In accordance with this definition, a “surgical process” can be defined as “network of surgical or surgery-related activities and their relationships, criteria to indicate the start and termination of the process, and the information about the individual activities”, whereas a “surgical work-step” or “activity” can be defined as “description of a piece of work that forms one logical step within a surgical process”.

Furthermore, a “model” is a simplified and constrained description of a fraction of reality that is built for a specific purpose. Thus, “modeling” can be understood as a simplified (formal or semiformal) representation of a suitable excerpt of reality using a suitable method for representation [24]. In consequence, an SPM can be defined as “a simplified (formal or semiformal) representation of a network of surgical or surgery-related activities and their relationships, criteria to indicate the start and termination of the process, and the information about the individual activities”. It is important to bear in mind that the simplified representation should reflect a predefined subset of interest [25] and that the method of representation should be suitable to the overall application of the model.

Delimitation

SPMs focus on the description of surgical work-steps. These activities represent the most important value-adding processes within the OR. Therefore, the goal of SPM needs to be to suitably represent these value-adding processes. Within an SPM, every surgical process needs to be represented as a chain or a network of single surgical or surgery-related work-steps (such as device setup in the OR, patient positioning, briefing, or instrument passing), including detailed information.

In most cases, SPMs cover the cut-suture-time of a surgical procedure. Activities before or after this period are usually not considered within SPMs. In general, SPMs focus on intraoperative work-steps. Perioperative events, such as anesthesiologic activities or activities performed by nonmedical staff within the OR, are of little or no importance.

However, there are some exceptions in which some of these external activities are of interest for the surgical workflow. This, for instance, is the case for a quantitative assessment of the impact of an SAS on the overall intervention course. In this case, the setup times of these systems, which occur preoperatively, need to be taken into consideration.

Model generation

This section will give an introduction on the strategies that can be employed for the generation of SPMs. Subsequently, strategies for the acquisition of data will be demonstrated. Third, the concept of model generalization will be explained. This strategy is mainly used for achieving an extended scope of application. Lastly, an overview on existing modeling languages will be given along with examples of use-cases.

Modeling strategies

In general, there are two different strategies for modeling surgical processes: top-down and bottom-up strategies.

Top-down modeling

This method can be employed if a process as a whole is too complex and must be broken down into single parts. The requirements for top-down modeling are simple but important. The modeling relies heavily on the interaction of experienced clinicians and process professionals. Top-down modeling results in a formulated overview of the process that is represented. Subsequently, every subpart of this process is refined and more details are added to the model. This method can be applied down to the desired level.

This approach offers some advantages. For instance, this modeling strategy is highly flexible and very readily available. Also, a breaking down of a procedure might help to better understand the single work-steps, simplifying each part of the process. In addition, parts of the models thus achieved can be reusable.

However, this approach also has some restrictions. For instance, this solution provides only limited coverage at the beginning. Furthermore, the possible resolution achieved by employing this method is rather limited and the degree of quantification is rather low, whereas the sheer number of work-steps represented can be overwhelming. Additionally, the experts using this approach to model a process might have problems with time estimation, rendering the models rather inaccurate. Also, the variability of the process might not be dutifully represented. Furthermore, there is a possibility that even an expert has wrong perceptions of certain aspects of the overall work, thus tainting the resulting model.

Current descriptions of surgical processes have further limitations. Sources, such as clinical guidelines or surgical textbooks, are valuable and easily accessible references to apply to the top-down modeling strategy. However, the applicability of these sources regarding analytical resolves has restrictions because of the deficient attention to detail, the lacking objective quantifiability, and the subjective point of view of the creator(s) of the model. Also, the variability of the process is insufficiently represented. Due to these facts, it is important and sensible to develop and implement new approaches for the acquisition of information about surgical processes, providing a basis to overcome the current limitations.

Bottom-up modeling

Bottom-up modeling works the other way around. In this approach, individual steps of a process are described and then linked together. These in turn are then interlinked until the overall process system is represented. The requirements for this strategy are an appropriate data acquisition modality and knowledge abstraction algorithms that give semantics to the big data.

The main disadvantage of this strategy is the very high complexity these models can achieve, rendering them very difficult to understand and process. The advantages, however, include a very high precision of the overall model, a possibility to consider process variability, and a possible acquisition of quantitative information, such as performance or usage time. Even if bottom-up modeling has some drawbacks, it is widely used.

Model types

Model information can be expressed by symbolic or numeric values. This section introduces both main types briefly. However, it is also possible to combine both model types. Top-down modeling results in symbolic model types, and bottom-up modeling can result in both types.

Symbolical models

Symbolical models are models that are expressed by means of natural language expression in plain text. The prerequisites of such models are a mutual terminology or the creation or definition of an underlying ontology. Furthermore, a sufficient computerized observation support for projects that need data acquisition during long time periods is highly recommended.

The advantages of these symbolical models are high flexibility and quick availability. Furthermore, there is no need to mine low-level semantics out of the data. Also, unexpected situations can be handled with relative ease while using this strategy. However, on the downside, the resolution of symbolical models is lower than that of sensor-based models. Furthermore, the risk of errors introduced by the observer is increased.

Numerical models

Numerical models are represented by physical measurements that are represented by numerical values, such as trajectories representing movements. The prerequisites for the application of numerical models are existing sensor systems and suitable knowledge abstraction algorithms to derive symbolic representations from the numerical models to verify it with clinicians.

The main advantage of numerical models is their very detailed level of representation. Additionally, due to the extraction of the data using sensors, acquisition errors due to human failure are relatively low. On the downside, semantic abstraction algorithms are needed and there are very few suitable sensor technologies available at the moment (see “Data acquisition by sensor systems”), although these technologies are rapidly ameliorated and researched. Additionally, these technologies are quite inflexible, can be quite expensive, and might need a high level of maintenance.

Data acquisition strategies for bottom-up modeling

There are different existing strategies for data acquisition. They all have in common that there is no “one and only” universal strategy to encompass the multitude of potentially interesting entities within the OR. In the following section, different data acquisition strategies will be described. As per definition, these strategies are not needed for top-down modeling strategies.

Data acquisition by observation

Data acquisition strategies are methods that are of interest to the process modeler and can be employed for the realization of surgical and technical use-cases. In [25], [26], a quick and flexible strategy for the acquisition of data produced by different entities in the OR by means of observation has been described. This strategy is based on the definition mentioned above, namely, that a surgical process is a succession of surgical work-steps [27], [28], [29]. These work-steps are in turn represented by various entities that are represented as perspectives. These perspectives are, for instance,

  • The functional perspective describing “what” is done in a surgical work-step,

  • The organizational perspective that describes “who” is performing a work-step,

  • The operational perspective that describes the “instrument(s)” used to perform a certain work-step,

  • The spatial perspective that represents “where” a workplace is performed (e.g. at which anatomical structure), and

  • The behavioral perspective that describes the “time” at which a certain work-step is performed.

Actual data acquisition is performed by means of tablet PCs that are equipped with a specialized software operated by especially trained observers (cf. Figure 1) [30]. The observers are present in the OR during the complete course of the intervention and record every surgical work-step they observe.

Figure 1: Workflow editor for recording SPMs (cataract surgery example configuration) [26].
Figure 1:

Workflow editor for recording SPMs (cataract surgery example configuration) [26].

This data acquisition strategy has been validated and it has been shown that the mean accuracy for the description of surgical work-step contents was approximately 92% and the temporal delay less than 2 s [25]. The benefit of this strategy is that it was unnecessary to include a work-step of knowledge abstraction in the processing pipeline that interprets numerical data into clinically interpretable knowledge, is usually surgery specific, and might be very complex.

Using this data acquisition strategy, more than 1000 single surgical processes have been recorded, thus demonstrating that the method can be used independently for different surgical disciplines. The recorded interventions were from ENT-surgery, cardiovascular surgery, neurosurgical surgery, pediatric surgery, and interventional radiology. The models thus acquired were of varying complexity according to the length and intricacy of the surgical process recorded. For instance, median model sizes for cataract surgeries comprised 20 activities. Other, more complex interventions comprised 100–150 activities (spinal disc surgery or sinus surgery) or 200–250 activities (mitral valve repair). The most complex interventions recorded, namely, interventions for brain tumor removal, comprised more than 400 single work-steps.

Data acquisition by sensor systems

Furthermore, there are many data acquisition strategies that rely on sensors rather than on human observers. In the past, several research groups using sensor systems have focused largely on a single perspective or have attempted to work without the use of an explicit process model. Additionally, many current works are centered on the usage of surgical instruments and device parameter recognition. The recognition of the actual intervention phase or state is a precondition for workflow analysis, optimization, and management.

The general challenge is that, in surgery, there usually are no intrinsic data or signal available to identify single manual tasks. To solve this problem, an analysis of endoscopic and microscopic video has been proposed [31], [32], [33]. Markov models and dynamic time warping (DTW) were used to identify single work-steps based on the presence of surgical instruments [34], [35], [36], [37], and radiofrequency identification (RFID), visual approaches, and weight analysis methods [38] have been employed to recognize instrument use [37], [39], [40]. For laparoscopy, another approach using a camera on the trocar and color wheels was proposed by Toti et al. [41]. Other technologies proposed were hand gesture recognition using thermal imaging [42] and accelerometer data analysis [43]. In robotic minimal invasive surgery (RMIS), a mixture of video and kinematic data was employed to identify gestures [44].

All of these approaches address different levels of granularity from ranged gesture recognition (surgemes) [42], [44], [45] to low-level tasks [33], high-level tasks [46], [47], and intervention phases [35], [48]. James et al. [49] tried to recognize the current surgical situation indirectly by estimating the positions and movements of the members of surgical teams within the OR or by deriving information from other indirect features. Most of these approaches achieve accuracies of between 80% and 90%. Additionally, this research, in general, focused on several different clinical use-cases, such as pituitary surgery [33], laparoscopic sigmoidectomy [50], cataract surgery [51], [52], or laparoscopic cholecystectomy [36].

Model generalization and transformation strategies

The ultimate goal for the developers of SPMs is the creation of models that can, according to the paradigm of individualization in Medicine 4.0, flexibly represent a multitude of surgical cases, disciplines, patients, and instruments. However, this aim is hard to achieve due to some characteristics of surgical processes, the most challenging being the high complexity and the high variability: adverse or unexpected turnouts of events are possible at every stage of the intervention, causing the overall workflow to deviate from the expected “standard”. This variability depends on many aspects, such as patient or surgeon properties, employed technical resources, and the hospital. Additionally, these interventions can be unstandardized as well as unplanned and otherwise highly variable.

To be able to encompass this high level of variability of surgical processes, approaches using top-down models, on the one hand, need to detail each possible outcome in sufficiently high resolution. Button-up model approaches, on the other hand, need to find strategies to transpose valid single recordings to represent other cases as well. This can be achieved by means of model generalization.

The basic idea of the generalization of workflow recordings is to merge activities. For an improved distinction, the terms patient-individual SPM (iSPM) will be used for an SPM before generalization and generic SPMs (gSPMs) will be used for an SPM after generalization subsequently.

The single process steps first need to be identified by defining certain features, such as, for instance, recognizing equal work-steps in different processes (cf. Figure 2). These work-steps can be used for process mining, a strategy transferred from business information systems. Thus, after the selection of the feature in focus, the various elements from the SPMs are merged and transferred to represent a new work-step in the generalized model. If a feature is unique, i.e. if it does not appear in any other recorded workflow, it is not merged and results in a new branch of the generalized SPM [53].

Figure 2: Simplified gSPM generation procedure example with surgical activities A …C and statistical transition probabilities.
Figure 2:

Simplified gSPM generation procedure example with surgical activities A …C and statistical transition probabilities.

This strategy can be supported by an application of sophisticated data mining methods [54]. The choice of these methods is dependent on the objective and the size of the actual study. It can be assumed that the higher the number of merged individual process models is, the higher and the more stable the resulting generalized process model will be. Figure 3 shows an example of a gSPM for activities of a surgeon while performing different surgical activities during cataract procedures.

Figure 3: gSPMs for Capsulorhexis phase in cataract interventions (solid line style, “mean” path of right hand; dotted line style, “mean” path of left hand).Nodes indicate activity descriptions and average performance time of the activity, and edge label indicate local (upper value, the predecessor-successor node probability related to the predecessor) and global transition probabilities (lower value, the predecessor-successor node probability related to the overall iSPM sample). The gSPM was filtered by removing edges with global transition probabilities of less than 5% to remove blur.
Figure 3:

gSPMs for Capsulorhexis phase in cataract interventions (solid line style, “mean” path of right hand; dotted line style, “mean” path of left hand).

Nodes indicate activity descriptions and average performance time of the activity, and edge label indicate local (upper value, the predecessor-successor node probability related to the predecessor) and global transition probabilities (lower value, the predecessor-successor node probability related to the overall iSPM sample). The gSPM was filtered by removing edges with global transition probabilities of less than 5% to remove blur.

In the example shown in Figure 3, 53 iSPMs of inpatient cataract surgeries were recorded and merged into a gSPM. The iSPMs containing surgical work-steps of the left and right hands of the surgeon were measured individually. Subsequently, activities of the same name in the iSPMs were combined into one activity in the gSPM, such as all activities of “paracentesis (with) paracentesis knife (with) right hand (at) cornea” (see bottom activity in Figure 3). If activities of the same name are followed by different activities, this results into a branching of the model. This approach enables different strategies: on the one hand, the gSPM contains all variability that was measured between the iSPMs; on the other hand, each iSPM is still present in the overall model and can be tracked to the gSPM.

Modeling languages

Basically, there is no designated or recommended language for the modeling of SPMs yet. Which language is chosen therefore entirely depends on the given circumstances, such as the requirements on the system, the scope of use of the models, the choice of system itself, or the expressive powers or the mightiness of the language.

In addition to the minimum requirement to be able to represent nodes and edges, the available elements of the language determine whether or not a language is applicable within a certain approach. Possible characteristics are the degree of mathematical formalization, the representation of spatial or temporal information, and the representation of probabilities but also the coverage of qualitative information such as the detection of conditions or risks associated with single work-steps.

A language with a high level of formalization should be employed whenever the process models to be represented need to be used as procedure algorithm for computer programs. The advantage of a language with a high formalization level furthermore is the possibility of verifying the models as early as during the design phase, for instance, with regard to the accessibility of work-steps or the presence of blind ends.

In general, there is a multitude of languages in existence that could be used to represent SPMs. Overall, a number of more than 650 process modeling languages are known. Most of these languages have their origin in information systems. The most influential and best-known of these languages for the field of business information systems are the XML Process Definition Language (XPDL) created by the Workflow Management Coalition (WFMC), the Business Process Modeling Notation (BPMN) and the Unified Modeling Language (UML) both developed by the Object Management Group (OMG), and the Business Process Execution Language (BPEL) created by the Organization for the Advancement of Structured Information Standards (OASIS). Other languages include the Event-Driven Process Chains (EPC) or Yet Another Workflow Language (YAWL). Languages with a higher level of formalization include Petri nets, state charts, and activity charts. A review and evaluation of modeling languages for SPM can be found in [55].

Model application examples

The application range of SPMs is diverse and multifaceted. In the following section, some possible applications will be presented, such as visualization, analysis, and workflow management.

SPMs for visualization and analysis

Alongside the performance of analyses, SPM is one of the easiest yet most effective application areas to allow for a useful and utilizable visualization of the course of a surgical intervention. In the past, it has become apparent that these visualized procedure courses are a useful support instrument for interdisciplinary work, such as a cooperation between surgeons and engineers. By means of this visualization, both professions can gain a thorough understanding of the single work-steps and the overall procedure, thus allowing them to combine their professional points of view to find trigger points for optimizations or to discuss the impact of new SAS.

The use of SPMs to the advantage of clinical users is a necessary criterion for an effective application. Moreover, the applicability across different intervention types and diverse disciplines of surgery is a key requirement. In the following section, the implementation of SPMs will be demonstrated in several use-cases.

Performing requirements analyses using SPMs

In [56], SPMs were used to derive certain working condition parameters for a neurological SAS from a number of iSPMs of the same procedure type. With the help of the parameters thus computed, the requirements for the operation of a milling system employing navigated-control to be used at the spine were projected. To realize this objective, neurosurgical interventions of lumbar discectomies were analyzed to predict time-based requirements for the computerized milling system.

The presented approach described an objective method for assessing the potential advantages of an SAS before its actual deployment or even before its development. The focus of this research was on the minimal requirements a projected system needs to be more advantageous than a conventional or existent method or system. Furthermore, the method to procure these requirements from routine clinical data was researched.

Lumbar discectomies are interventions in which surgeons extract material from the lumbar disc(s) of a patient. First, the surgeon removes a part of the vertebra to gain access to the actual location of the material to be extracted. Second, the surgeon removes as much of the lumbar disk as might be reached. These two steps are iterated until the targeted material has been completely extracted. This strategy is employed to minimize the impact on the vertebra and its supporting function.

The research question in this study was whether an alternative strategy employing an automated milling system was more effective than the conventional method. The difference between the conventional system and the new system was that, in the latter, the mill worked automatically within a predefined volume, namely, the total amount of material to be extracted. The new computer-assisted system was expected to eliminate the need of iterations, thus shortening the procedure to the benefit of patients, hospital, and OR staff.

Human observers recorded the interventions with the help of software tools shown in Figure 1. These recordings were computed into gSPMs. Subsequently, the relevant work-steps (e.g. “removing disk tissue”) were analyzed according to their distribution along the procedure (e.g. the work-step iteration of removing tissue and removing bone material was performed ~15 times during the procedure) and their temporal occurrence. The latter revealed that the automated milling system should not perform longer than ~20 min per procedure [56]. The information was than processed to calculate how fast the new system needed to work to gain a quantitative advantage on the conventional method. This was mainly possible thanks to the detailed and exact-to-the-second recording and representation of surgical activities.

The methods presented in this work are transferrable to other procedure types and might be used for a systematic analysis of other requirements as well. For instance, it can be used for a systematic analysis of technical requirements. Based on such analyses, manufacturers can derive parameters for planned systems that are indicative of their success or failure on the market. Additionally, hospital administrations might request that manufacturers demonstrate fulfillment of such requirements to prove the superiority of their system and the worth of the investment.

With regard to the overall goals of improving and planning novel systems based on SPMs, this approach method creates a better and more appropriate support systems for the surgeon by integrating the surgical workflow into development considerations.

Evaluation of SAS using SPMs

The question of whether a new SAS contributes to the surgical performance needs to be answered qualitatively and quantitatively, supported by evidence gathered from structured data analyses. In [57], gSPMs were used to evaluate SAS. Laparoscopic and telemanipulator-based Nissen fundoplications in pediatric surgery were compared, with the aim of calculating the impact of the telemanipulator on the overall process. In the study, the system’s impact on the process was investigated and quantified based on process models. By comparing both gSPMs for the laparoscopic and telemanipulator-based strategies, it could be demonstrated that the use of a telemanipulator is not recommended for this clinical use-case. This approach demonstrated to be a valuable method to estimate the impact of an SAS on the surgical workflow. The SPM analysis results demonstrated that the telemanipulator strategy followed to close the laparoscopic strategy. Hence, significant faster intervention times could not be achieved even for subphases or work-steps [57] or even were less efficient due to the spatial and sensual limitations. For the investigated use-case, changes in the overall course of the intervention may be necessary to fully benefit from the advantages of a telemanipulator in Nissen fundoplications.

Evaluation and comparison of surgical approaches

In [58], SPMs were used to optimize patient treatment strategies. In this study, a potential advanced training for surgeons from an ophthalmological department of a university medical center was planned based on the models generated from their performed interventions, called surgeon-individual treatment profiles (SiTPs). The surgical interventions observed were standard cataract procedures. These were performed by three experienced surgeons during their daily routine. A total of 105 procedures were recorded and computed used to generate iSPMs. For each surgeon, separately, the resulting iSPMs were again computed into gSPMs (cf. Figure 4). Surgeon-individual workflows, activity frequencies, and median performance durations of activities were analyzed and compared. By comparing the three SiTPs, concrete recommendations for further training could be derived. Especially for gSPM segments with extended performance times or an increased number of iterations of surgical work-steps, the respective surgeon might benefit from an individualized and specific training.

Figure 4: SiTPs as gSPMs of three surgeons for the intervention phase Capsulorhexis.The graph shows the most commonly recorded work-steps for surgeons “red”, “blue”, and “green”. The activities are labeled in the nodes and the edges are labeled with global predecessor-successor probabilities. The gSPM was filtered to remove transitions with low probabilities to remove blurring.
Figure 4:

SiTPs as gSPMs of three surgeons for the intervention phase Capsulorhexis.

The graph shows the most commonly recorded work-steps for surgeons “red”, “blue”, and “green”. The activities are labeled in the nodes and the edges are labeled with global predecessor-successor probabilities. The gSPM was filtered to remove transitions with low probabilities to remove blurring.

Process compliance assessment

Until now, the use for SPMs to perform requirements analyses, evaluations, and comparison has been presented. In [59], these models were used to assess the quality of the outcome of surgical interventions using process benchmarking. This approach examined the compliance of 450 surgical processes assessing the relationship between the course of the intervention and its outcome.

For this aim, skill practices using rapid prototyping models in minimally invasive surgery training were assessed. From these assessments, representations of surgical processes were extracted and compared by applying process-oriented similarity metrics to the outcome of the “best practice” procedure. Thus, it was shown that a high process compliance that corresponds to low process deviation significantly supports good intervention outcomes, which is relevant for optimal patient care. Also, this method can be used to give the surgeons feedback with regard to human factors and also for inducing further changes in the overall workflow.

SPMs for workflow management in the OR

In modern ORs, the need for intelligent information presentation has been identified to prevent an information overload for surgeons and other OR staff [60], [61], [62], [63]. In the pertaining literature, context-dependent support is frequently mentioned as one of the chief objectives in the research field of SPMs [64], [65], [66], [67], [68]. However, these systems need to be intelligible and not overtaxing for their user(s) [69] as well as suitably balanced between proactivity and transparency [70].

Proactive support is based on appropriate SPMs. In [67], [71], the design and implementation of a surgical workflow management system based on SPMs that can provide a robust guidance for surgical activities was described. The number of required iSPMs was investigated by a randomized selection of iSPMs to create a gSPM and subsequent testing of a disjunct iSPM against the gSPM. Additionally, the method of investigating the number of SPMs needed to develop such a system was investigated. The models were mapped onto workflow nets, a dialect of Petri nets, as workflow execution templates to outline the behavior of the intraoperative workflow management system. The application of SPMs for the implementation of situation-dependent assistance is expected to reduce the workload of the surgeon during endoscopically performed ENT procedures as described in [72], [73], [74].

There are many possible application fields of surgical workflow management systems within the scope of Medicine 4.0 in the digital OR of the future with integrated medical device networks. These include programmed parameterization and governing of SAS, a situation- and context-dependent information visualization for surgeons, resulting, for instance, in a timely presentation of previously acquired patient examination results or the providing of decision support systems for resident surgeons. Additionally, the ensuing surgical workflow management system could be linked to the hospital information system (HIS), thus allowing for an automatic and timely call for the next patient according to the predicted beginning of the next intervention. These use-cases could ameliorate the surgical process qualitatively and therefore be beneficial to the patients’ safety.

The developing integration of modern medical devices and IT systems in the ORs of the future is an empowering technology. A wide variety of implementations of SAS – from management and administration to support – may be constructed based on data provided by medical devices, such as online prediction of the remaining procedure time [75] or estimating the probability of medical device applications [68].

Conclusion

The ensuing synergy between surgeons and technology is not yet fully matured, leading to breaches in the surgical workflow, delays of operating times, and media disruptions. Additionally, only fragments of the “big data” accrued in ORs by surgeons and machines are not used to its full extent. This is mainly due to infrastructure breaches and communication errors. The ICT used in hospitals cannot manage the clinical and operational sequences effectively.

To improve this state, SPMs, defined as simplified, formal, or semiformal representations of a network of surgical or surgery-related activities and their relationships, reflecting a predefined subset of interest, can be used. To create such models, various strategies can be used such as top-down or bottom-up modeling, different underlying languages or ontologies might be employed, and different data acquisition strategies can be used such as benchmarking and data mining. As a result, these models can represent surgical interventions down to a very fine granularity level in very high resolution, thus contributing qualifiable and quantifiable information on the intervention course, on interventional phases, or on surgical work-steps.

The focus of this research lies on the gathering of information concerning surgical intervention and its underlying activities. Such information includes the use of surgical tools, task duration, trajectories of instruments or movements of persons in the OR, or intervention phase recognition. The methods to accrue these data are manifold, such as a manual or automatic workflow recording, knowledge abstraction, live or camera recordings. These data are then abstracted to represent the surgical process as a model. Further abstraction can be generated by merging several iSPMs to create more gSPMs that are valid for larger patient samples.

The resulting SPMs can be used in a widespread variety of use-cases in the context of Medicine 4.0. Examples are requirements analyses, evaluating SAS, generating surgeon-specific training-recommendation, creating workflow management systems for ORs, and comparing different surgical strategies. In the future, the research field of SPMs will become more and more important for the orchestration of technology and for a situation-dependent support of the surgeon. It will be employed to create and manage the advances that are brought about due to arising new medical technologies and growing knowledge. Additionally, there is also a high probability that the research field of SPMs will spawn a technology of its own.

Author Statement

  1. Research funding: The research was funded by the German Ministry of Education and Research. Conflict of interest: Authors state no conflict of interest. Informed consent: Informed consent is not applicable. Ethical approval: The conducted research is not related to either human or animal use.

Author Contributions

  1. Thomas Neumuth: Writing of the manuscript; Approval of the manuscript.

Publication Funding

  1. The German Society of Surgery funded the article processing charges of this article.

References

[1] Stitzenberg KB, Sheldon GF. Progressive specialization within general surgery: adding to the complexity of workforce planning. J Am Coll Surg 2005;201:925–932.10.1016/j.jamcollsurg.2005.06.253Search in Google Scholar PubMed

[2] Dexter F, Abouleish AE, Epstein RH, Whitten CW, Lubarsky DA. Use of operating room information system data to predict the impact of reducing turnover times on staffing costs. Anesth Analg 2003;97: 1119–1126.10.1213/01.ANE.0000082520.68800.79Search in Google Scholar PubMed

[3] Dexter F, Epstein RH, Traub RD, Xiao Y. Making management decisions on the day of surgery based on operating room efficiency and patient waiting times. J Am Soc Anesthesiol 2004;101:1444–1453.10.1097/00000542-200412000-00027Search in Google Scholar PubMed

[4] Friedman DM, Sokal SM, Chang Y, Berger DL. Increasing operating room efficiency through parallel processing. Ann Surg 2006;243:10–14.10.1097/01.sla.0000193600.97748.b1Search in Google Scholar PubMed PubMed Central

[5] Kreyher VJ. Gesundheits-und Medizinmarketing-Herausforderung für das Gesundheitswesen. Handbuch Gesundheits-und Medizinmarketing. Chancen, Strategien und Erfolgsfaktoren, Heidelberg; 2001:1–52.Search in Google Scholar

[6] Stahl JE, Egan MT, Goldman JM, et al. Introducing new technology into the operating room: measuring the impact on job performance and satisfaction. Surgery 2005;137:518–526.10.1016/j.surg.2004.12.015Search in Google Scholar PubMed

[7] Baumgart A, Schüpfer G, Welker A, Bender H-J, Schleppers A. Status quo and current trends of operating room management in Germany. Curr Opin Anesthesiol 2010;23:193–200.10.1097/ACO.0b013e328336b8b4Search in Google Scholar PubMed

[8] Archer T, Macario A. The drive for operating room efficiency will increase quality of patient care. Curr Opin Anesthesiol 2006;19:171–176.10.1097/01.aco.0000192796.02797.82Search in Google Scholar PubMed

[9] Geldner G, Eberhart LHJ, Trunk S, et al. Effizientes OP-Management Vorschläge zur Optimierung von Prozessabläufen als Grundlage für die Erstellung eines OP-Statuts. Anaesthesist 2002;51:760–767.10.1007/s00101-002-0362-1Search in Google Scholar PubMed

[10] Sandberg WS, Ganous TJ, Steiner C. Setting a research agenda for perioperative systems design. Surg Innov 2003;10:57–70.10.1177/107155170301000202Search in Google Scholar PubMed

[11] Sunyaev A, Leimeister JM, Schweiger A, Krcmar H. Integrationsarchitekturen für das Krankenhaus. IMC Inf Manag Consult 2006;21:28–35.Search in Google Scholar

[12] Mauro C, Leimeister JM, Krcmar H. Serviceorientierte Integration medizinischer Geräte-ganzheitliche IT-Unterstützung klinischer Prozesse. Inf Spekt 2011;34:276–285.10.1007/s00287-010-0436-4Search in Google Scholar

[13] Documet J, Le A, Liu B, Chiu J, Huang H. A multimedia electronic patient record (ePR) system for image-assisted minimally invasive spinal surgery. Int J Comput Assist Radiol Surg 2010;5:195–209.10.1007/s11548-009-0387-xSearch in Google Scholar PubMed PubMed Central

[14] Gadatsch A. Management von Geschäftsprozessen. Methoden und Werkzeuge für die IT-Praxis, Braunschweig/Wiesbaden; 2002.10.1007/978-3-322-91983-0Search in Google Scholar

[15] Rosenkranz F. Geschäftsprozesse: modell-und computergestützte Planung. Springer-Verlag; 2005.Search in Google Scholar

[16] Scheer A-W. Wirtschaftsinformatik Studienausgabe: Referenzmodelle für industrielle Geschäftsprozesse. Springer-Verlag; 2013.Search in Google Scholar

[17] Scheer A-W. ARIS – Modellierungsmethoden, Metamodelle, Anwendungen. Springer-Verlag; 2013.Search in Google Scholar

[18] Neumuth T. Surgical Process Modeling –– Theory, Methods, and Applications. Habilitationsschrift, Universität Leipzig, Leipzig; 2013.Search in Google Scholar

[19] Oxford English Dictionary. 2016.Search in Google Scholar

[20] Merriam-Webster. Surgery | Definition of Surgery. 2016.Search in Google Scholar

[21] Surgery, Cambridge Dictionary. 2016.Search in Google Scholar

[22] German Ministry for Education and Research, Zukunftsbild Industrie 4.0. German Ministry for Education and Research; 2013.Search in Google Scholar

[23] WfMC, Workflow Management Coalition Terminology and Glossary (WFMC-TC-1011) | BibSonomy. Workflow Management Coalition; 1999.Search in Google Scholar

[24] Bungartz H-J. Modellbildung und Simulation Eine anwendungsorientierte Einführung 2., überarbeitete Auflage. Berlin, Heidelberg: Springer Spektrum; 2013.10.1007/978-3-642-37656-6_1Search in Google Scholar

[25] Neumuth T, Jannin P, Strauss G, Meixensberger J, Burgert O. Validation of knowledge acquisition for surgical process models. J Am Med Inform Assoc 2009;16:72–80.10.1197/jamia.M2748Search in Google Scholar PubMed PubMed Central

[26] Neumuth T, Kaschek B, Neumuth D, et al. An observation support system with an adaptive ontology-driven user interface for the modeling of complex behaviors during surgical interventions. Behav Res Methods 2010;42:1049–1058.10.3758/BRM.42.4.1049Search in Google Scholar PubMed

[27] Neumuth D, Loebe F, Herre H, Neumuth T. Modeling surgical processes: a four-level translational approach. Artif Intell Med 2011;51:147–161.10.1016/j.artmed.2010.12.003Search in Google Scholar PubMed

[28] Mudunuri R, Burgert O, Neumuth T. Ontological modelling of surgical knowledge. In: Fischer S, Maehle E, Reischuk R, editors. INFORMATIK 2009 – Im Focus das Leben, GI-Edition Lecture Notes in Informatics. Lübeck: Bonner Köllen Verlag; 2009:P-154.Search in Google Scholar

[29] Neumann J, Schreiber E, Neumuth T. Ontology-based surgical process modeling by using SNOMED CT concepts and concept model attributes. In: 30th International Conference on Computer Assisted Radiology and Surgery (CARS), Heidelberg, Germany, vol. 11;2016:S66–S68.Search in Google Scholar

[30] Neumuth T, Mudunuri R, Jannin P, Meixensberger J, Burgert O. SWAN-Suite: the tool landscape for surgical workflow analysis. In: Troccaz J, Merloz P, editors. Computer Assisted Medical and Surgical Interventions. Paris: Sauchamps Medical; 2007:199–204.Search in Google Scholar

[31] Bhatia B, Oates T, Xiao Y, Hu P. Real-time identification of operating room state from video. In: Proceedings of the 19th National Conference on Innovative Applications of Artificial Intelligence – Volume 2, Vancouver, British Columbia, Canada; 2007:1761–1766.Search in Google Scholar

[32] Speidel S, Benzko J, Krappe S, et al. Automatic classification of minimally invasive instruments based on endoscopic image sequences. Proc. SPIE 7261, Medical Imaging 2009: Visualization, Image-Guided Procedures, and Modeling 2009; 72610A.10.1117/12.811112Search in Google Scholar

[33] Lalys F, Riffaud L, Morandi X, Jannin P. Automatic phases recognition in pituitary surgeries by microscope images classification. In: Navab N, Jannin P, editors. Information Processing in Computer-Assisted Interventions. Berlin/Heidelberg: Springer; 2010:34–44.10.1007/978-3-642-13711-2_4Search in Google Scholar

[34] Rosen J, Brown JD, Chang L, Sinanan MN, Hannaford B. Generalized approach for modeling minimally invasive surgery as a stochastic process using a discrete Markov model. IEEE Trans Biomed Eng 2006;53:399–413.10.1109/TBME.2005.869771Search in Google Scholar PubMed

[35] Blum T, Padoy N, Feußner H, Navab N. Modeling and online recognition of surgical phases using hidden Markov models. In: Metaxas D, Axel L, Fichtinger G, Székely G, editors. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2008. Berlin/Heidelberg: Springer; 2008:627–635.10.1007/978-3-540-85990-1_75Search in Google Scholar PubMed

[36] Padoy N, Blum T, Ahmadi S-A, Feussner H, Berger M-O, Navab N. Statistical modeling and recognition of surgical workflow. Med Image Anal 2012;16:632–641.10.1016/j.media.2010.10.001Search in Google Scholar PubMed

[37] Meißner C, Neumuth T. RFID-based surgical instrument detection using hidden Markov models. Biomed Eng/Biomed Tech 57, no. SI-1 Track-H; 2012.10.1515/bmt-2012-4047Search in Google Scholar

[38] Glaser B, Dänzer S, Neumuth T. Intra-operative surgical instrument usage detection on a multi-sensor table. Int J Comput Assist Radiol Surg 2015;10:351–362.10.1007/s11548-014-1066-0Search in Google Scholar PubMed

[39] Kranzfelder M, Schneider A, Fiolka A, et al. Real-time instrument detection in minimally invasive surgery using radiofrequency identification technology. J Surg Res 2013;185:704–710.10.1016/j.jss.2013.06.022Search in Google Scholar PubMed

[40] Meißner C, Meixensberger J, Pretschner A, Neumuth T. Sensor-based surgical activity recognition in unconstrained environments. Minim Invasive Ther Allied Technol 2014;23:198–205.10.3109/13645706.2013.878363Search in Google Scholar PubMed

[41] Toti G, Garbey M, Sherman V, Bass BL, Dunkin BJ. A smart trocar for automatic tool recognition in laparoscopic surgery. Surg Innov 2015;22:77–82.10.1177/1553350614531659Search in Google Scholar PubMed

[42] Unger M, Chalopin C, Neumuth T. Vision-based online recognition of surgical activities. Int J Comput Assist Radiol Surg 2014;9:979–986.10.1007/s11548-014-0994-zSearch in Google Scholar PubMed

[43] Ahmadi S-A, Padoy N, Rybachuk K, Feussner H, Heinin S, Navab N. Motif discovery in OR sensor data with application to surgical workflow analysis and activity detection. In: M2CAI Workshop, MICCAI, London; 2009.Search in Google Scholar

[44] Zappella L, Béjar B, Hager G, Vidal R. Surgical gesture classification from video and kinematic data. Med Image Anal 2013;17:732–745.10.1016/j.media.2013.04.007Search in Google Scholar PubMed

[45] Tao L, Zappella L, Hager GD, Vidal R. Surgical gesture segmentation and recognition. In: Mori K, Sakuma I, Sato Y, Barillot C, Navab N, editors. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2013. Berlin/Heidelberg: Springer; 2013:339–346.10.1007/978-3-642-40760-4_43Search in Google Scholar PubMed

[46] Lalys F, Riffaud L, Bouget D, Jannin P. An application-dependent framework for the recognition of high-level surgical tasks in the OR. In: Fichtinger G, Martel A, Peters T, editors. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2011. Berlin/Heidelberg: Springer; 2011:331–338.10.1007/978-3-642-23623-5_42Search in Google Scholar PubMed

[47] Bouarfa L, Jonker PP, Dankelman J. Discovery of high-level tasks in the operating room. J Biomed Inform 2011;44: 455–462.10.1016/j.jbi.2010.01.004Search in Google Scholar PubMed

[48] Stauder R, Okur A, Peter L, et al. Random forests for phase detection in surgical workflow analysis. In: Stoyanov D, Collins DL, Sakuma I, Abolmaesumi P, Jannin P, editors. Information Processing in Computer-Assisted Interventions 8498. Cham: Springer International Publishing; 2014:148–157.10.1007/978-3-319-07521-1_16Search in Google Scholar

[49] James A, Vieira D, Lo B, Darzi A, Yang GZ. Eye-gaze driven surgical workflow segmentation. Med Image Comput Comput Assist Interv 2007;10:110–117.10.1007/978-3-540-75759-7_14Search in Google Scholar PubMed

[50] Weede O, Dittrich F, Worn H, et al. Workflow analysis and surgical phase recognition in minimally invasive surgery. In: 2012 IEEE International Conference on Robotics and Biomimetics (ROBIO);2012:1080–1074.10.1109/ROBIO.2012.6491111Search in Google Scholar

[51] Lalys F, Bouget D, Riffaud L, Jannin P. Automatic knowledge-based recognition of low-level tasks in ophthalmological procedures. Int J Comput Assist Radiol Surg 2012;8:39–49.10.1007/s11548-012-0685-6Search in Google Scholar PubMed

[52] Quellec G, Charrière K, Lamard M, et al. Real-time recognition of surgical tasks in eye surgery videos. Med Image Anal 2014;18:579–590.10.1016/j.media.2014.02.007Search in Google Scholar PubMed

[53] Neumuth T, Jannin P, Schlomberg J, Meixensberger J, Wiedemann P, Burgert O. Analysis of surgical intervention populations using generic surgical process models. Int J Comput Assist Radiol Surg 2011;6:59–71.10.1007/s11548-010-0475-ySearch in Google Scholar PubMed PubMed Central

[54] van der Aalst W. Business alignment: using process mining as a tool for delta analysis and conformance testing. Require Eng 2005;10:198–211.10.1007/s00766-005-0001-xSearch in Google Scholar

[55] Neumann J, Rockstroh M, Vinz S, Neumuth T. Technical Report: Surgical Workflow and Process Modeling; 2015.Search in Google Scholar

[56] Neumuth T, Trantakis C, Riffaud L, Strauss G, Meixensberger J, Burgert O. Assessment of technical needs for surgical equipment by surgical process models, Minimally Invasive Ther Allied Technol 2009;18:341–349.10.3109/13645700903384484Search in Google Scholar PubMed

[57] Neumuth T, Krauss A, Meixensberger J, Muensterer OJ. Impact quantification of the daVinci telemanipulator system on surgical workflow using resource impact profiles, Int J Med Robot Comput Assist Surg 2011;7:156–164.10.1002/rcs.383Search in Google Scholar PubMed

[58] Neumuth T, Wiedemann R, Foja C, et al. Identification of surgeon-individual treatment profiles to support the provision of an optimum treatment service for cataract patients. J Ocul Biol Dis Inform 2010;3:73–83.10.1007/s12177-011-9058-6Search in Google Scholar PubMed PubMed Central

[59] Schumann S, Bühligen U, Neumuth T. Outcome quality assessment by surgical process compliance measures in laparoscopic surgery. Artif Intell Med 2015;63:85–90.10.1016/j.artmed.2014.10.008Search in Google Scholar PubMed

[60] Pickering BW, Herasevich V, Ahmed A, Gajic O. Novel representation of clinical information in the ICU: developing user interfaces which reduce information overload. Appl Clin Inform 2010;1:116–131.10.4338/ACI-2009-12-CR-0027Search in Google Scholar PubMed PubMed Central

[61] Rockstroh M, Franke S, Neumuth T. A workflow-driven surgical information source management. Int J Comput Assist Radiol Surg 2013;8:189–191.Search in Google Scholar

[62] Koninckx PR, Stepanian A, Adamyan L, Ussia A, Donnez J, Wattiez A. The digital operating room and the surgeon. Gynecol Surg 2013;10:57–62.10.1007/s10397-012-0752-6Search in Google Scholar

[63] Kranzfelder M, Staub C, Fiolka A, et al. Toward increased autonomy in the surgical OR: needs, requests, and expectations. Surg Endosc 2013;27:1681–1688.10.1007/s00464-012-2656-ySearch in Google Scholar PubMed

[64] Speidel S, Zentek T, Sudra G, et al. Recognition of surgical skills using hidden Markov models. SPIE Proc 2009;7261:726125–726125-8.10.1117/12.811140Search in Google Scholar

[65] Katić D, Wekerle A-L, Gärtnet F,et al. Knowledge-driven formalization of laparoscopic surgeries for rule-based intraoperative context-aware assistance. In: Stoyanov D, Collins DL, Sakuma I, Abolmaesumi P, Jannin P, editors. Information Processing in Computer-Assisted Interventions 8498. Cham: Springer International Publishing; 2014:158–167.10.1007/978-3-319-07521-1_17Search in Google Scholar

[66] Kranzfelder M, Schneider A, Fiolka A, et al. Reliability of sensor-based real-time workflow recognition in laparoscopic cholecystectomy. Int J Comput Assist Radiol Surg 2014;9:941–948.10.1007/s11548-014-0986-zSearch in Google Scholar PubMed

[67] Franke S, Neumuth T. A framework for event-driven surgical workflow assistance. Biomed Eng/Biomed Tech 2014;59:431–434.Search in Google Scholar

[68] Franke S, Meixensberger J, Neumuth T. Multi-perspective workflow modeling for online surgical situation models. J Biomed Inform 2015;54:158–166.10.1016/j.jbi.2015.02.005Search in Google Scholar PubMed

[69] Bricon-Souf N, Newman CR. Context awareness in health care: a review. Int J Med Inform 2007;76:2–12.10.1016/j.ijmedinf.2006.01.003Search in Google Scholar PubMed

[70] Satyanarayanan M. Pervasive computing: vision and challenges. IEEE Pers Commun 2001;8:10–17.10.1109/98.943998Search in Google Scholar

[71] Franke S, Liebmann P, Neumuth T. Connecting workflow management to the OR network: design and evaluation of a bridge to enable dynamic systems behaviour. Biomed Tech 2012;57:1.10.1515/bmt-2012-4192Search in Google Scholar PubMed

[72] Franke S, Neumuth T. Rule-based medical device adaptation for the digital operating room: In Proceedings of 37th Annual International Conference of the IEEE Engineering in Medicine and Biology Society, Milano; 2015:1733–1736.10.1109/EMBC.2015.7318712Search in Google Scholar PubMed

[73] Franke S, Rockstroh M, Schreiber E, Neumann J, Neumuth T. Towards the intelligent OR – implementation of distributed, context-aware automation in an integrated surgical working environment. Proc. of the 7th Workshop on Modeling and Monitoring of Computer Assisted Interventions (M2CAI) at the 19th International Conference on Medical Image Computing and Computer Assisted Interventions (MICCAI); 2016. In press.Search in Google Scholar

[74] Franke S, Rockstroh M, Schreiber E, Neumann J, Neumuth T. Context-aware medical assistance systems in integrated surgical environments. In: 28th Conference of the International Society for Medical Innovation and Technology (SMIT); 2016. In press.Search in Google Scholar

[75] Franke S, Meixensberger J, Neumuth T. Intervention time prediction from surgical low-level tasks. J Biomed Inform 2013;46: 152–159.10.1016/j.jbi.2012.10.002Search in Google Scholar PubMed


Supplementary Material:

The article (DOI: https://doi.org/10.1515/iss-2017-0005) offers reviewer assessments as supplementary material.


Received: 2017-01-17
Accepted: 2017-04-05
Published Online: 2017-05-20

©2017 Neumuth T., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 License.

Downloaded on 5.6.2024 from https://www.degruyter.com/document/doi/10.1515/iss-2017-0005/html
Scroll to top button